Download The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Biochemical switches in the cell cycle wikipedia , lookup

Thylakoid wikipedia , lookup

Cell encapsulation wikipedia , lookup

Cellular differentiation wikipedia , lookup

Extracellular matrix wikipedia , lookup

Cell culture wikipedia , lookup

Programmed cell death wikipedia , lookup

Cell growth wikipedia , lookup

Cell cycle wikipedia , lookup

Organ-on-a-chip wikipedia , lookup

Signal transduction wikipedia , lookup

Cytosol wikipedia , lookup

Cell membrane wikipedia , lookup

Cytoplasmic streaming wikipedia , lookup

Cell nucleus wikipedia , lookup

JADE1 wikipedia , lookup

Cytokinesis wikipedia , lookup

Mitosis wikipedia , lookup

Amitosis wikipedia , lookup

Endomembrane system wikipedia , lookup

List of types of proteins wikipedia , lookup

Transcript
CHAPTER TWO
The Protoplast: Plasma Membrane,
Nucleus, and Cytoplasmic Organelles
Cells represent the smallest structural and functional
units of life (Sitte, 1992). Living organisms consist of
single cells or of complexes of cells. Cells vary greatly
in size, form, structure, and function. Some are measured in micrometers, others in millimeters, and still
others in centimeters (fibers in certain plants). Some
cells perform a number of functions; others are specialized in their activities. Despite the extraordinary diversity among cells they are remarkably similar to one
another, both in their physical organization and in their
biochemical properties.
The concept that the cell is the basic unit of biological structure and function is based on the cell theory,
which was formulated in the first half of the nineteenth
century by Mathias Schleiden and Theodor Schwann. In
1838, Schleiden concluded that all plant tissues are composed of cells. A year later, Schwann (1839) extended
Schleiden’s observation to animal tissues and proposed
a cellular basis for all life. In 1858, the idea that all living
organisms are composed of one or more cells took on
even broader significance when Rudolf Virchow generalized that all cells arise only from preexisting cells. In
its classical form, the cell theory proposed that the
bodies of all plants and animals are aggregates of individual, differentiated cells, and that the activities of the
whole plant and animal might be considered the summation of the activities of the individual constituent
cells, with the individual cells of prime importance.
By the latter half of the nineteenth century, an alternative to the cell theory was formulated. Known as the
organismal theory, it maintains that the entire organism is not merely a group of independent units but
rather a living unit subdivided into cells, which are connected and coordinated into a harmonious whole. An
often quoted statement is that of Anton de Bary (1879),
“It is the plant that forms cells, and not the cell that
forms plants” (translation by Sitte, 1992). Since then
substantial evidence has accumulated in favor of
an organismal concept for plants (see Kaplan and
Hagemann, 1991; Cooke and Lu, 1992; and Kaplan,
1992; and literature cited therein).
The organismal theory is especially applicable to
plants, whose cells do not pinch apart during cell division, as do animal cells, but are partitioned initially by
insertion of a cell plate (Chapter 4). The separation of
plant cells is rarely complete. Contiguous plant cells
Esau’s Plant Anatomy, Third Edition, By Ray F. Evert.
Copyright © 2006 John Wiley & Sons, Inc.
15
16 | Esau’s Plant Anatomy, Third Edition
remain interconnected by cytoplasmic strands known
as plasmodesmata, which traverse the walls and unite
the entire plant body into an organic whole. Appropriately, plants have been characterized as supracellular
organisms (Lucas et al., 1993).
In its modern form the cell theory states simply that:
(1) all organisms are composed of one or more cells, (2)
the chemical reactions of a living organism, including
its energy-related processes and its biosynthetic processes, occur within cells, (3) cells arise from other
cells, and (4) cells contain the hereditary information of
the organisms of which they are a part, and this information is passed on from parent to daughter cell. The cell
and organismal theories are not mutually exclusive.
Together, they provide a meaningful view of the structure and function at cellular and organismal levels (Sitte,
1992).
The word cell, meaning “little room,” was introduced
by Robert Hooke in the seventeenth century to describe
the small cavities separated by cell walls in cork tissue.
Later Hooke recognized that living cells in other plant
tissues were fi lled with “juices.” Eventually the contents
of cells were interpreted as living matter and received
the name protoplasm. An important step toward recognition of the complexity of protoplasm was the discovery of the nucleus by Robert Brown in 1831. This
discovery was soon followed by reports of cell division.
In 1846, Hugo von Mohl called attention to the distinction between protoplasmic material and cell sap, and in
1862, Albert von Kölliker used the term cytoplasm for
the material surrounding the nucleus. The most conspicuous inclusions in the cytoplasm, the plastids, were
long considered to be merely condensations of protoplasm. The concept of independent identity and continuity of these organelles was established in the
nineteenth century. In 1880, Johannes Hanstein introduced the term protoplast to designate the unit of
protoplasm inside the cell wall.
Every living cell has a means of isolating its contents
from the external environment. A membrane called the
plasma membrane, or plasmalemma, brings about
this isolation. Plant cells have, in addition, a more or less
rigid cellulosic cell wall (Chapter 4) deposited outside
the plasma membrane. The plasma membrane controls
the passage of materials into and out of the protoplast
and so makes it possible for the cell to differ structurally
and biochemically from its surroundings. Processes
within a cell can release and transfer the energy necessary for growth and for the maintenance of metabolic
processes. A cell is organized to retain and transfer
information so that its development and that of its
progeny can occur in an orderly manner. This way the
integrity of the organism, of which the cells are a part,
is maintained.
In the three centuries since Hooke first observed the
structure of cork through his rudimentary microscope,
our capacity to see the cell and its contents has increased
dramatically. With improvement of the light microscope,
it became possible to observe objects with a diameter
of 0.2 micrometer (about 200 nanometers), an improvement on the naked eye about 500 times. With the transmission electron microscope (TEM), the limit of
resolution imposed by light was greatly reduced. Because
of problems with specimen preparation, contrast, and
radiation damage, however, the resolution of biological
objects is more like 2 nanometers. Nonetheless, this is
still 100 times better than the resolution of the light
microscope. The TEM has distinct disadvantages,
however: the specimen to be observed must be preserved (dead) and cut into exceedingly thin, effectively
two-dimensional slices. Optical microscopy using fluorescent dyes and various methods of illumination have
enabled biologists to overcome these problems and to
observe subcellular components in live plant cells
(Fricker and Oparka, 1999; Cutler and Ehrhardt, 2000).
Notable is the use of green fluorescent protein (GFP),
from the jelly fish Aequorea victoria, as a fluorescent
protein tag and of confocal microscopy to visualize the
fluorescent probes in intact tissues (Hepler and Gunning,
1998; Fricker and Oparka, 1999; Hawes et al., 2001). The
observation of subcellular components in live plant cells
is providing new and often unexpected insights into
subcellular organization and dynamics.
❙ PROKARYOTIC AND EUKARYOTIC CELLS
Based on the degree of internal organization of their
cells, two fundamentally distinct groups of organisms
are now recognized: prokaryotes and eukaryotes. The
prokaryotes (pro, before; karyon, nucleus) are represented by the Archaea and Bacteria, including the cyanobacteria, and the eukaryotes (eu, true; karyon, nucleus)
by all other living organisms (Madigan et al., 2003).
Prokaryotic cells differ most notably from eukaryotic
cells in the organization of their genetic material. In
prokaryotic cells, the genetic material is in the form of
a large, circular molecule of deoxyribonucleic acid
(DNA), with which a variety of proteins are loosely
associated. This molecule, which is called the bacterial
chromosome, is localized in a region of the cytoplasm
called the nucleoid (Fig. 2.1). In eukaryotic cells, the
nuclear DNA is linear and tightly bound to special proteins known as histones, forming a number of more
complex chromosomes. These chromosomes are surrounded by a nuclear envelope, made up of two membranes, that separates them from the other cellular
contents in a distinct nucleus (Fig. 2.2). Both prokaryotic cells and eukaryotic cells contain complexes of
protein and ribonucleic acid (RNA), known as ribosomes, that play a crucial role in the assembly of protein
molecules from their amino acid subunits.
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 17
DNA
region
plasma
membrane
ribosomes
0.2 mm
FIGURE 2.1
Electron micrograph of the gram-negative bacterium,
Azotobacter vinelandii. The granular appearance of
the cytoplasm is largely due to the presence of numerous ribosomes. The clearer DNA-containing regions
constitute the nucleoid. (Courtesy of Jack L. Pate.)
Eukaryotic cells are subdivided by membranes into
distinct compartments that perform different functions.
The cytoplasm of prokaryotic cells, by contrast, typically is not compartmentalized by membranes. Notable
exceptions are the extensive system of photosynthetic
membranes (thylakoids) of the cyanobacteria (Madigan
et al., 2003) and the membrane-bounded entities called
acidocalcisomes found in a variety of bacteria, including
Agrobacterium tumefaciens, the plant pathogen that
causes crown gall (Seufferheld et al., 2003).
The appearance of membranes under the electron
microscope is remarkably similar in various organisms.
When suitably preserved and stained, these membranes
have a three-layered appearance, consisting of two dark
layers separated by a lighter layer (Fig. 2.3). This type
of membrane was named unit membrane by
Robertson (1962) and interpreted as a bimolecular lipid
layer covered on each side with a layer of protein.
Although this model of membrane structure has been
superseded by the fluid mosaic model (see below), the
term unit membrane remains a useful designation for a
visually definable three-ply membrane.
Among the internal membranes of eukaryotic cells
are those surrounding the nucleus, mitochondria, and
plastids, which are characteristic components of plant
cells. The cytoplasm of eukaryotic cells also contains
systems of membranes (the endoplasmic reticulum and
Golgi apparatus) and a complex network of nonmembranous protein fi laments (actin fi laments and microtubules) called the cytoskeleton. A cytoskeleton is absent
in prokaryotic cells. Plant cells also develop multifunctional organelles, called vacuoles, that are bound by a
membrane called the tonoplast (Fig. 2.2).
In addition to the plasma membrane, which controls
the passage of substances into and out of the protoplast,
the internal membranes control the passage of substances among compartments within the cell. This way
the cell can maintain the specialized chemical environments necessary for the processes occurring in the different cytoplasmic compartments. Membranes also
permit differences in electrical potential, or voltage, to
become established between the cell and its environment and between adjacent compartments of the cell.
Differences in the chemical concentration of various
ions and molecules and the electric potential across
membranes provide potential energy used to power
many cellular processes.
Compartmentation of cellular contents means division of labor at the subcellular level. In a multicellular
organism a division of labor occurs also at the cellular
level as the cells differentiate and become more or less
specialized with reference to certain functions. Functional specialization finds its expression in morphological differences among cells, a feature that accounts
for the complexity of structure in a multicellular
organism.
❙ CYTOPLASM
As mentioned previously, the term cytoplasm was
introduced to designate the protoplasmic material surrounding the nucleus. In time, discrete entities were
discovered in this material, first only those that were
within the resolving power of the light microscope;
later, smaller entities were discovered with the electron
microscope. Thus the concept of cytoplasm has undergone an evolution; with new technologies the concept
undoubtedly will continue to evolve. Most biologists
today use the term cytoplasm, as originally introduced
by Kölliker (1862), to designate all of the material surrounding the nucleus, and they refer to the cytoplasmic
18 | Esau’s Plant Anatomy, Third Edition
er
w
nu
nu
n
p
m
o
ne
p
v
o
v
5 mm
FIGURE 2.2
Nicotiana tabacum (tobacco) root tip. Longitudinal section of young cells. Details: er, endoplasmic reticulum; m,
mitochondrion; n, nucleus; ne, nuclear envelope; nu, nucleolus; o, oil body; p, plastid; v, vacuole; w, cell wall. (From
Esau, 1977.)
matrix, in which the nucleus, organelles, membrane
systems, and nonmembranous entities are suspended,
as the cytosol. As originally defi ned, however, the term
cytosol was used to refer specifically “to the cytoplasm
minus mitochondria and endoplasmic reticulum components” in liver cells (Lardy, 1965). Cytoplasmic ground
substance and hyaloplasm are terms that commonly
have been used by plant cytologists to refer to the cytoplasmic matrix. Some biologists use the term cytoplasm
in the sense of cytosol.
In the living plant cell the cytoplasm is always in
motion; the organelles and other entities suspended in
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 19
TABLE 2.1 ■ An Inventory of Plant Cell Components
mt
mt
Cell wall
Protoplast
pm
Middle lamella
Primary wall
Secondary wall
Plasmodesmata
Nucleus
Cytoplasm
cell wall
pm
mt
mt
mt
0.12 mm
FIGURE 2.3
Electron micrograph showing the three-layered appearance of the plasma membranes (pm) on either side of
the common wall between two cells of an Allium cepa
leaf. Microtubules (mt) in transectional view can be
seen on both sides of the wall.
the cytosol can be observed being swept along in an
orderly fashion in the moving currents. This movement,
which is known as cytoplasmic streaming, or cyclosis, results from an interaction between bundles of actin
fi laments and the so-called motor protein, myosin, a
protein molecule with an ATPase-containing “head” that
is activated by actin (Baskin, 2000; Reichelt and
Kendrich-Jones, 2000). Cytoplasmic streaming, a costly
energy-consuming process, undoubtedly facilitates the
exchange of materials within the cell (Reuzeau et al.,
1997; Kost and Chua, 2002) and between the cell and
its environment.
The various components of the protoplast are considered individually in the following paragraphs. Among
those components are the entities called organelles. As
with the term cytoplasm, the term organelle is used differently by different biologists. Whereas some restrict use
of the term organelle to membrane-bound entities such
as plastids and mitochondria, others use the term more
broadly to refer also to the endoplasmic reticulum and
Golgi bodies and to nonmembranous components such
as microtubules and ribosomes. The term organelle is
used in the restricted sense in this book (Table 2.1).
Nuclear envelope
Nucleoplasm
Chromatin
Nucleolus
Plasma membrane
Cytosol (cytoplasmic
ground substance,
hyaloplasm)
Organelles bounded by
two membranes:
Plastids
Mitochondria
Organelles bounded
by one membrane:
Peroxisomes
Vacuoles, bounded
by tonoplast
Ribosomes
Endomembrane system
(major components):
Endoplasmic reticulum
Golgi apparatus
Vesicles
Cytoskeleton:
Microtubules
Actin filaments
In this chapter only the plasma membrane, nucleus, and
cytoplasmic organelles are considered. The remaining
components of the protoplast are covered in Chapter 3.
❙ PLASMA MEMBRANE
Among the various membranes of the cell, the plasma
membrane typically has the clearest dark-light-dark or
unit membrane appearance in electron micrographs
(Fig. 2.3; Leonard and Hodges, 1980; Robinson, 1985).
The plasma membrane has several important functions:
(1) it mediates the transport of substances into and out
of the protoplast, (2) it coordinates the synthesis and
assembly of cell wall microfibrils (cellulose), and (3) it
transduces hormonal and environmental signals involved
in the control of cell growth and differentiation.
The plasma membrane has the same basic structure
as the internal membranes of the cell, consisting of a
lipid bilayer in which are embedded globular proteins,
many extending across the bilayer and protrude on
either side (Fig. 2.4). The portion of these transmembrane proteins embedded in the bilayer is
20 | Esau’s Plant Anatomy, Third Edition
carbohydrate
outside of cell
lipid
bilayer
peripheral
protein
hydrophobic
zone
hydrophilic
zone
inside of cell
FIGURE 2.4
Fluid-mosaic model of membrane structure. The membrane is composed of a bilayer of lipid molecules—with their
hydrophobic “tails” facing inward—and large protein molecules. Some of the proteins (transmembrane proteins)
traverse the bilayer; others (peripheral proteins) are attached to the transmembrane proteins. Short carbohydrate
chains are attached to most of the protruding transmembrane proteins on the outer surface of the plasma membrane.
The whole structure is quite fluid; some of the transmembrane proteins float freely within the bilayer, and together
with the lipid molecules move laterally within it, forming different patterns, or “mosaics,” and hence the proteins
can be thought of as floating in a lipid “sea.” (From Raven et al., 1992.)
hydrophobic, whereas the portion or portions exposed
on either side of the membrane are hydrophilic.
The inner and outer surfaces of a membrane differ
considerably in chemical composition. For example,
there are two major types of lipids in the plasma membrane of plant cells—phospholipids (the more abundant) and sterols (particularly stigmasterol)—and the
two layers of the bilayer have different compositions of
these. Moreover the transmembrane proteins have definite orientations within the bilayer, and the portions
protruding on either side have different amino acid
compositions and tertiary structures. Other proteins are
also associated with membranes, including the peripheral proteins, so called because they lack discrete
hydrophobic sequences and thus do not penetrate into
the lipid bilayer. Transmembrane proteins and other
lipid-bound proteins tightly bound to the membrane are
called integral proteins. On the outer surface of the
plasma membrane, short-chain carbohydrates (oligosaccharides) are attached to the protruding proteins,
forming glycoproteins. The carbohydrates, which form
a coat on the outer surface of the membranes of some
eukaryotic cells, are believed to play important roles in
cell-to-cell adhesion processes and in the “recognition”
of molecules (e.g., hormones, viruses, and antibiotics)
that interact with the cell.
Whereas the lipid bilayer provides the basic structure
and impermeable nature of cellular membranes, the proteins are responsible for most membrane functions.
Most membranes are composed of 40% to 50% lipid (by
weight) and 60% to 50% protein, but the amounts and
types of proteins in a membrane reflect its function.
Membranes involved with energy transduction, such as
the internal membranes of mitochondria and chloroplasts, consist of about 75% protein. Some of the proteins are enzymes that catalyze membrane-associated
reactions, whereas others are transport proteins
involved in the transfer of specific molecules into and
out of the cell or organelle. Still others act as receptors
for receiving and transducing chemical signals from the
cell’s internal or external environment. Although some
of the integral proteins appear to be anchored in place
(perhaps to the cytoskeleton), the lipid bilayer is generally quite fluid. Some of the proteins float more or less
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 21
freely in the bilayer, and they and the lipid molecules
can move laterally within it, forming different patterns,
or mosaics, that vary from time to time and place to
place—hence the name fluid-mosaic for this model of
membrane structure (Fig. 2.4; Singer and Nicolson,
1972; Jacobson et al., 1995).
Membranes contain different kinds of transport proteins (Logan et al., 1997; Chrispeels et al., 1999;
Kjellbom et al., 1999; Delrot et al., 2001). Two of the
types are carrier proteins and channel proteins, both of
which permit the movement of a substance across a
membrane only down the substance’s electrochemical
gradient; that is, they are passive transporters. Carrier
proteins bind the specific solute being transported and
undergo a series of conformational changes in order to
transport the solute across the membrane. Channel
proteins form water-fi lled pores that extend across the
membrane and, when open, allow specific solutes
(usually inorganic ions, e.g., K + , Na + , Ca2+ , Cl−) to pass
through them. The channels are not open continuously;
instead they have “gates” that open briefly and then
close again, a process referred to as gating.
The plasma membrane and tonoplast also contain
water channel proteins called aquaporins that specifically facilitate the passage of water through the
membranes (Schäffner, 1998; Chrispeels et al., 1999;
Maeshima, 2001; Javot and Maurel, 2002). Water passes
relatively freely across the lipid bilayer of biological
membranes, but the aquaporins allow water to diffuse
more rapidly across the plasma membrane and tonoplast. Because the vacuole and cytosol must be in constant osmotic equilibrium, rapid movement of water is
essential. It has been suggested that aquaporins facilitate the rapid flow of water from the soil into root cells
and to the xylem during periods of high transpiration.
Aquaporins have been shown to block the influx of
water into cells of the roots during periods of flooding
(Tournaire-Roux et al., 2003) and to play a role in
drought avoidance in rice (Lian et al., 2004). In addition
evidence indicates that water movement through aquaporins increases in response to certain environmental
stimuli that induce cell expansion and growth; the
cyclic expression of a plasma membrane aquaporin has
been implicated in the leaf unfolding mechanism in
tobacco (Siefritz et al., 2004).
Carriers can be classified as uniporters and cotransporters according to how they function. Uniporters
transport only one solute from one side of the membrane to another. With cotransporters, the transfer of
one solute depends on the simultaneous or sequential
transfer of a second solute. The second solute may be
transported in the same direction, in which case the
carrier protein is known as symporter, or in the opposite direction, as in the case of an antiporter.
The transport of a substance against its electrochemical gradient requires the input of energy, and is called
active transport. In plants that energy is provided
primarily by an ATP-powered proton pump, specifically, a membrane-bound H + -ATPase (Sze et al., 1999;
Palmgren, 2001). The enzyme generates a large gradient
of protons (H + ions) across the membrane. This gradient
provides the driving force for solute uptake by all protoncoupled cotransport systems. The tonoplast is unique
among plant membranes in having two proton pumps,
an H + -ATPase and an H + -pyrophosphatase (H + -PPase)
(Maeshima, 2001), although some data indicate that H + PPase may also be present in the plasma membrane of
some tissues (Ratajczak et al., 1999; Maeshima, 2001).
The transport of large molecules such as most proteins and polysaccharides cannot be accommodated by
the transport proteins that ferry ions and small polar
molecules across the plasma membrane. These large
molecules are transported by means of vesicles or saclike structures that bud off from or fuse with the plasma
membrane, a process called vesicle-mediated transport (Battey et al., 1999). The transport of material into
the cell by vesicles that bud off of the plasma membrane
is called endocytosis and involves portions of the
plasma membrane called coated pits (Fig. 2.5; Robinson
A
0.1 mm
C
B
0.1 mm
0.1 mm
FIGURE 2.5
Endocytosis in maize (Zea mays) rootcap cells that have
been exposed to a solution containing lead nitrate. A,
granular deposits containing lead can be seen in two
coated pits. B, a coated vesicle with lead deposits. C,
here, one of two coated vesicles has fused with a large
Golgi vesicle where it will release its contents. This
coated vesicle (dark structure) still contains lead deposits, but it appears to have lost its coat, which is located
just to the right of it. The coated vesicle to its left is
clearly intact. (Courtesy of David G. Robinson.)
22 | Esau’s Plant Anatomy, Third Edition
and Depta, 1988; Gaidarov et al., 1999). Coated pits are
depressions in the plasma membrane containing specific receptors (to which the molecules to be transported into the cell must first bind) and coated on their
cytoplasmic surface with clathrin, a protein composed
of three large and three smaller polypeptide chains that
together form a three-pronged structure, called a triskelion. Invaginations of the coated pits pinch off to form
coated vesicles. Within the cell the coated vesicles
shed their coats and then fuse with some other membrane-bound structures (e.g., Golgi bodies or small vacuoles). Transport by means of vesicles in the opposite
direction is called exocytosis (Battey et al., 1999).
During exocytosis, vesicles originating from within the
cell fuse with the plasma membrane, expelling their
contents to the outside.
Relatively large invaginations, or infoldings, of the
plasma membrane are frequently encountered in tissue
prepared for electron microscopy. Some form pockets
between the cell wall and protoplast, and may include
tubules and vesicles. Some invaginations may push the
tonoplast forward and intrude into the vacuole. Others,
called multivesicular bodies, are often detached from
the plasma membrane and embedded in the cytosol or
appear suspended in the vacuole. Similar formations
were first observed in fungi and named lomasomes
(Clowes and Juniper, 1968). Multivesicular bodies in
Nicotiana tabacum BY-2 cells have been identified as
plant prevacuolar compartments that lie on the endocytic pathway to lytic vacuoles (see below; Tse et al.,
2004).
the presence of a great many cylindrical nuclear pores,
which provide direct contact between the cytosol and
the ground substance, or nucleoplasm, of the nucleus
(Fig. 2.6). The inner and outer membranes are joined
around each pore, forming the margin of its opening.
Structurally complicated nuclear pore complexes—
the largest supramolecular complexes assembled in the
eukaryotic cell—span the envelope at the nuclear pores
(Heese-Peck and Raikhel, 1998; Talcott and Moore,
1999; Lee, J.-Y., et al., 2000). The nuclear pore complex
is roughly wheel-shaped, consisting in part of a cylindrical central channel (the hub) from which eight spokes
project outwardly to an interlocking collar associated
with the nuclear membrane lining the pore. The nuclear
pore complexes allow relatively free passage of certain
ions and small molecules through diffusion channels,
which measure about 9 nanometers in diameter. The
proteins and other macromolecules transported through
the nuclear pore complexes greatly exceed this channel
size. Their transport is mediated by a highly selective
ribosomes
po
❙ NUCLEUS
Often the most prominent structure within the protoplast of eukaryotic cells, the nucleus performs two
important functions: (1) it controls the ongoing activities of the cell by determining which RNA and protein
molecules are produced by the cell and when they are
produced, and (2) it is the repository of most of the cell’s
genetic information, passing it on to the daughter cells
in the course of cell division. The total genetic information stored in the nucleus is referred to as the nuclear
genome.
The nucleus is bounded by a pair of membranes
called the nuclear envelope, with a perinuclear
space between them (Figs. 2.2 and 2.6; Dingwall and
Laskey, 1992; Gerace and Foisner, 1994; Gant and Wilson,
1997; Rose et al., 2004). In various places the outer
membrane of the envelope is continuous with the endoplasmic reticulum, so that the perinuclear space is continuous with the lumen of the endoplasmic reticulum.
The nuclear envelope is considered a specialized, locally
differentiated portion of the endoplasmic reticulum.
The most distinctive feature of the nuclear envelope is
annulus
ne
po
B
microtubule
0.5 mm
A
FIGURE 2.6
Nuclear envelope (ne) in profi le (A) and from the surface
(B, central part) showing pores (po). The electron-dense
material in the pores in A is shown, in B, to have a form
of an annulus with a central granule. The clear space
between the membranes in A is called the perinuclear
space. From a parenchyma cell in Mimosa pudica
petiole. (From Esau, 1977.)
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 23
active (energy-dependent) transport mechanism that
takes place through the central channel. The central
channel has a functional diameter of up to 26 nanometers (Hicks and Raikhel, 1995; Görlich and Mattaj, 1996;
Görlich, 1997).
In specially stained cells, thin threads and grains of
chromatin can be distinguished from the nucleoplasm.
Chromatin is made up of DNA combined with large
amounts of proteins called histones. During the process
of nuclear division, the chromatin becomes progressively more condensed until it takes the form of chromosomes. Chromosomes (chromatin) of nondividing,
or interphase, nuclei are attached at one or more sites
to the inner membrane of the nuclear envelope. Before
DNA replication each chromosome is composed of a
single, long DNA molecule, which carries the hereditary
information. In most interphase nuclei the bulk of chromatin is diffuse and lightly staining. This uncondensed
chromatin, called euchromatin, is genetically active
and associated with high rates of RNA synthesis. The
remaining, condensed chromatin, called heterochromatin, is genetically inactive; that is, it is not associated
with RNA synthesis (Franklin and Cande, 1999). Overall,
only a small percentage of the total chromosomal DNA
codes for essential proteins or RNAs; apparently there
is a substantial surplus of DNA in the genomes of higher
organisms (Price, 1988). Nuclei may contain proteinaceous inclusions of unknown function in crystalline,
fibrous, or amorphous form (Wergin et al., 1970), in
addition to chromatin-containing “micropuffs” and
coiled bodies composed of ribonucleoprotein (Martín
et al., 1992).
Different organisms vary in the number of chromosomes present in their somatic (vegetative, or body)
cells. Haplopappus gracilis, a desert annual, has 4 chromosomes per cell; Arabidopsis thaliana, 10; Vicia faba,
broad bean, 12; Brassica oleracea, cabbage, 18; Asparagus officinalis, 20; Triticum vulgare, bread wheat, 42;
and Cucurbita maxima, squash, 48. The reproductive
cells, or gametes, have only half the number of chromosomes that is characteristic of the somatic cells in an
organism. The number of chromosomes in the gametes
is referred to as the haploid (single set) number and
designated as n, and that in the somatic cells is called
the diploid (double set) number, which is designated
as 2n. Cells that have more than two sets of chromosomes are said to be polyploid (3n, 4n, 5n, or more).
Often the only structures discernible within a nucleus
with the light microscope are spherical structures
known as nucleoli (singular: nucleolus) (Fig. 2.2;
Scheer et al., 1993). The nucleolus contains high concentrations of RNA and proteins, along with large loops
of DNA emanating from several chromosomes. The
loops of DNA, known as nucleolar organizer regions,
contain clusters of ribosomal RNA (rRNA) genes. At
these sites, newly formed rRNAs are packaged with
ribosomal proteins imported from the cytosol to form
ribosomal subunits (large and small). The ribosomal
subunits are then transferred, via the nuclear pores, to
the cytosol where they are assembled to form ribosomes. Although the nucleolus commonly is thought of
as the site of ribosome manufacture, it is involved with
only a part of the process. The very presence of a nucleolus is due to the accumulation of the molecules being
packaged to form ribosomal subunits.
In many diploid organisms, the nucleus contains one
nucleolus to each haploid set of chromosomes. The
nucleoli may fuse and then appear as one large structure. The size of a nucleolus is a reflection of the level
of its activity. In addition to the DNA of the nucleolar
organizer region, nucleoli contain a fibrillar component
consisting of rRNA already associated with protein to
form fibrils, and a granular component consisting of
maturing ribosomal subunits. Active nucleoli also show
lightly stained regions commonly referred to as vacuoles. In living cultured cells these regions, which should
not be confused with the membrane-bound vacuoles
found in the cytosol, can be seen to be undergoing
repeated contractions, a phenomenon that might be
involved with RNA transport.
Nuclear divisions are of two kinds: mitosis, during
which a nucleus gives rise to two daughter nuclei, each
morphologically and genetically equivalent to the other
and to the parent nucleus; meiosis, during which the
parent nucleus undergoes two divisions, one of which
is a reduction division. By a precise mechanism, meiosis
produces four daughter nuclei, each with one-half the
number of chromosomes as the parent nucleus. In
plants, mitosis gives rise to somatic cells and to gametes
(sperm and egg), and meiosis to meiospores. In both
kinds of division (with some exceptions) the nuclear
envelope breaks into fragments, which become indistinguishable from ER cisternae, and the nuclear pore complexes are disassembled. When new nuclei are assembled
during telophase, ER vesicles join to form two nuclear
envelopes, and new nuclear pore complexes are formed
(Gerace and Foisner, 1994). The nucleoli disperse during
late prophase (with some exceptions) and are newly
organized during telophase.
❙ CELL CYCLE
Actively dividing somatic cells pass through a regular
sequence of events known as the cell cycle. The cell
cycle commonly is divided into interphase and mitosis
(Fig. 2.7; Strange, 1992). Interphase precedes mitosis,
and in most cells, mitosis is followed by cytokinesis,
the division of the cytoplasmic portion of a cell and the
separation of daughter nuclei into separate cells (Chapter
4). Hence most plant cells are uninucleate. Certain specialized cells may become multinucleate either only
24 | Esau’s Plant Anatomy, Third Edition
G2 phase: Structures
required for cell division
begin to assemble;
chromosomes begin to
G2 checkpoint
condense.
M phase: The two sets
of chromosomes are
separated (mitosis)
and the cell divides
(cytokinesis).
M
Division
G2
G1
S
Interphase
S phase: DNA replicated
and associated proteins
synthesized; two copies of
cell’s genetic information
now exist.
G1 phase: Cell doubles
in size; organelles,
enzymes, and other
molecules increase
in number.
G1 checkpoint
FIGURE 2.7
The cell cycle. Cell division, which consists of mitosis (the division of the nucleus) and cytokinesis (the division of
the cytoplasm), takes place after the completion of the three preparatory phases (G1, S, and G2) of interphase. Progression of the cell cycle is mainly controlled at two checkpoints, one at the end of G1 and the other at the end of
G2. After the G2 phase comes mitosis, which is usually followed by cytokinesis. Together, mitosis and cytokinesis
constitute the M phase of the cell cycle. In cells of different species or of different tissues within the same organism,
the various phases occupy different proportions of the total cycle. (From Raven et al., 2005.)
during their development (e.g., nuclear endosperm) or
for life (e.g., nonarticulated laticifers). Mitosis and cytokinesis together are referred to as the M phase of the
cell cycle.
Interphase can be divided into three phases, which
are designated G1, S, and G2. The G1 phase (G stands for
gap) occurs after mitosis. It is a period of intense biochemical activity, during which the cell increases in
size, and the various organelles, internal membranes,
and other cytoplasmic components increase in number.
The S (synthesis) phase is the period of DNA replication. At the onset of DNA replication, a diploid nucleus
is said to have a 2C DNA value (C is the haploid DNA
content); at completion of the S phase, the DNA value
has doubled to 4C. During the S phase, many of the
histones and other DNA-associated proteins are also synthesized. Following the S phase, the cell enters the G2
phase, which follows the S phase and precedes mitosis.
The primary role of the S phase is to make sure chromosome replication is complete and to allow for repair of
damaged DNA. The microtubules of the preprophase
band, a ring-like band of microtubules that borders the
plasma membrane and encircles the nucleus in a plane
corresponding to the plane of cell division, also develop
during the G2 phase (Chapter 4; Gunning and Sammut,
1990). During mitosis the genetic material synthesized
during the S phase is divided equally between two
daughter nuclei, restoring the 2C DNA value.
The nature of the control or controls that regulate
the cell cycle apparently is basically similar in all eukaryotic cells. In the typical cell cycle, progression is
controlled at crucial transition points, called checkpoints—first at the G1-S phase transition and then at the
G2-M transition (Boniotti and Griffith, 2002). The first
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 25
checkpoint determines whether or not the cell enters
the S phase, the second whether or not mitosis is initiated. A third checkpoint, the metaphase checkpoint,
delays anaphase if some chromosomes are not properly
attached to the mitotic spindle. Progression through the
cycle depends on the successful formation, activation,
and subsequent inactivation of cyclin-dependent protein
kinases (CDKs) at the checkpoints. These kinases consist
of a catalytic CDK subunit and an activating cyclin
subunit (Hemerly et al., 1999; Huntley and Murray, 1999;
Mironov et al., 1999; Potuschak and Doerner, 2001; Stals
and Inzé, 2001). Both auxins and cytokinins have been
implicated in the control of the plant cell cycle
(Jacqmard et al., 1994; Ivanova and Rost, 1998; den Boer
and Murray, 2000).
Cells in the G1 phase have several options. In the
presence of sufficient stimuli they can commit to further
cell division and progress into the S phase. They may
pause in their progress through the cell cycle in response
to environmental factors, as during winter dormancy,
and resume dividing at a later time. This specialized
resting, or dormant, state is often called the Go phase
(G-zero phase). Other fates include differentiation and
programmed cell death, a genetically determined
program that can orchestrate death of the cell (Chapter
5; Lam et al., 1999).
Some cells feature only DNA replication and gap
phases without subsequent nuclear division, a process
known as endoreduplication (Chapter 5; D’Amato,
1998; Larkins et al., 2001). The single nucleus then
becomes polyploid (endopolyploidy, or endoploidy).
Endoploidy may be part of the differentiation of single
cells, as it is in the Arabidopsis trichome (Chapter 9),
or that of any tissue or organ. A positive correlation
exists between cell volume and the degree of polyploidy
in most plant cells, indicating that polyploid nuclei
might be required for the formation of large plant cells
(Kondorosi et al., 2000).
❙ PLASTIDS
Together with vacuoles and cell walls, plastids are characteristic components of plant cells (Bowsher and Tobin,
2001). Each plastid is surrounded by an envelope consisting of two membranes. Internally the plastid is differentiated into a more or less homogeneous matrix, the
stroma, and a system of membranes called thylakoids.
The principal permeability barrier between cytosol and
plastid stroma is the inner membrane of the plastid
envelope. The outer membrane, although a barrier to
cytosolic proteins, has generally been assumed to be
permeable to low molecular weight solutes (<600 Da),
an assumption that may be in need of re-evaluation
(Bölter and Soll, 2001). Stroma-filled tubules have been
observed emanating from the surfaces of some plastids.
These so-called stromules can interconnect different
plastids and have been shown to permit exchange of
green fluorescent protein between plastids (Köhler
et al., 1997; Köhler and Hanson, 2000; Arimura et al.,
2001; Gray et al., 2001; Pyke and Howells, 2002; Kwok
and Hanson, 2004). In a study of stromule biogenesis,
increases in stromule length and frequency correlated
with chromoplast differentiation; it was proposed that
stromules enhance the specific metabolic activities of
plastids (Waters et al., 2004).
Plastids are semiautonomous organelles widely accepted to have evolved from free-living cyanobacteria
through the process of endosymbiosis (Palmer and
Delwiche, 1998; Martin, 1999; McFadden, 1999;
Reumann and Keegstra, 1999; Stoebe and Maier, 2002).
Indeed, plastids resemble bacteria in several ways. For
example, plastids, like bacteria, contain nucleoids,
which are regions containing DNA. The DNA of the
plastid, like that of the bacterium, exists in circular form
(Sugiura, 1989); moreover it is not associated with histones. During the course of evolution most of the DNA
of the endosymbiont (the cyanobacterium) was gradually transferred to the host nucleus; hence the genome
of the modern plastid is quite small compared to the
nuclear genome (Bruce, 2000; Rujan and Martin, 2001).
Both plastids and bacteria contain ribosomes (70S ribosomes) that are about two-thirds as large as the ribosomes (80S ribosomes) found in the cytosol and
associated with endoplasmic reticulum. (The S stands
for Svedbergs, the units of the sedimentation coefficient.) In addition the process of plastid division—binary
fission—is morphologically similar to bacterial cell
division.
Chloroplasts Contain Chlorophyll and
Carotenoid Pigments
Mature plastids are commonly classified on the basis of
the kinds of pigments they contain. Chloroplasts (Figs.
2.8–2.10), the sites of photosynthesis, contain chlorophyll and carotenoid pigments. The chlorophyll pigments are responsible for the green color of these
plastids, which occur in green plant parts and are particularly numerous and well differentiated in leaves. In
seed plants, chloroplasts are usually disk-shaped and
measure between 4 and 6 micrometers in diameter. The
number of chloroplasts found in a single mesophyll
(middle of the leaf) cell varies widely, depending on the
species and the size of the cell (Gray, 1996). A single
mesophyll cell of cocoa (Cacao theobroma) and Peperomia metallia leaves may contain as few as three
chloroplasts, whereas as many as 300 chloroplasts occur
in a single mesophyll cell of the radish (Raphanus
sativus) leaf. The mesophyll cells of most leaves that
have been examined for plastid development contain 50
to 150 chloroplasts each. The chloroplasts are usually
26 | Esau’s Plant Anatomy, Third Edition
intermembrane space
stroma
thylakoids
granum
thylakoid
granum
outer membrane
stroma
inner membrane
FIGURE 2.8
Three-dimensional structure of a chloroplast. Note that
the internal membranes (thylakoids) are not connected
with the plastid envelope. (From Raven et al., 1992.)
found with their broad surfaces parallel to the cell wall,
preferentially on cell surfaces bordering air spaces.
They can reorient in the cell under the influence of
light—for example, gathering along the walls parallel
with the leaf surface under low or medium light intensity, thereby optimizing light utilization for photosynthesis (Trojan and Gabryś, 1996; Williams et al., 2003).
Under potentially damaging high light intensity the
chloroplasts can orient themselves along walls perpendicular to the leaf surface. The blue-UV region of the
spectrum is the most effective stimulus for chloroplast
movement (Trojan and Gabryś, 1996; Yatsuhashi, 1996;
Kagawa and Wada, 2000, 2002). In the darkness the
chloroplasts are distributed either randomly around all
the cell walls or their arrangement depends on local
factors inside the cells (Haupt and Scheuerlein, 1990).
Presumably movement of the chloroplasts involves an
actin-myosin-based system.
The internal structure of the chloroplast is complex.
The stroma is traversed by an elaborate system of thylakoids, consisting of grana (singular: granum)—stacks
of disk-like thylakoids that resemble a stack of coins—
and stroma thylakoids (or intergrana thylakoids) that
traverse the stroma between grana and interconnect
them (Figs. 2.8–2.10). The grana and stroma thylakoids
and their internal compartments are believed to constitute a single, interconnected system. The thylakoids are
not physically connected with the plastid envelope but
are completely embedded in the stroma. Chlorophylls
and carotenoid pigments—both of which are involved
in light harvesting—are embedded, along with proteins,
in the thylakoid membranes in discrete units of organization called photosystems. The principal function of
the carotenoid pigments is that of an antioxidant,
preventing photo-oxidative damage to the chlorophyll
molecules (Cunningham and Gantt, 1998; Vishnevetsky
et al., 1999; Niyogi, 2000).
Chloroplasts often contain starch, phytoferritin (an
iron compound) and lipid in the form of globules called
plastoglobuli (singular: plastoglobule). The starch
grains are temporary storage products and accumulate
only when the plant is actively photosynthesizing. They
may be lacking in the chloroplasts of plants kept in the
dark for as little as 24 hours but often reappear after the
plant has been in the light for only 3 or 4 hours.
Mature chloroplasts contain numerous copies of a
circular plastid DNA molecule and the machinery for
the replication, transcription, and translation of that
genetic material (Gray, J. C., 1996). With the limited
coding capacity (approximately 100 proteins) of the
chloroplast, however, the vast majority of proteins
involved with chloroplast biogenesis and function are
encoded by the nuclear genome (Fulgosi and Soll, 2001).
These proteins, which are synthesized on ribosomes in
the cytosol, are targeted into the chloroplast as precursor proteins with the aid of an amino-terminal extension referred to as a transit peptide. Each protein
imported into the chloroplast contains a specific transit
peptide. The transit peptide both targets the protein to
the chloroplast and mediates import into the stroma
where it is cleaved off after import (Flügge, 1990;
Smeekens et al., 1990; Theg and Scott, 1993). Transport
across a thylakoid membrane is mediated by a second
transit peptide unmasked when the first one is cleaved
off (Cline et al., 1993; Keegstra and Cline, 1999). Evidence indicates that part of the chloroplastic protein
machinery is derived from the endosymbiotic cyanobacterial ancestor of chloroplasts (Reumann and Keegstra,
1999; Bruce, 2000).
In addition to regulatory traffic from the nucleus to
the chloroplast, the chloroplasts transmit signals to the
nucleus to coordinate nuclear and chloroplast gene
expression. Moreover plastid signals also regulate the
expression of nuclear genes for nonplastid proteins and
for the expression of mitochondrial genes (see references in Rodermel, 2001). Chloroplasts are not only
sites of photosynthesis; they are also involved in amino
acid synthesis and fatty acid synthesis and provide space
for the temporary storage of starch.
Chromoplasts Contain Only Carotenoid Pigments
Chromoplasts (chroma, color) are also pigmented
plastids (Fig. 2.11). Of variable shape, they lack chlorophyll but synthesize and retain carotenoid pigments,
which are often responsible for the yellow, orange, or
red colors of many flowers, old leaves, some fruits, and
some roots. Chromoplasts are the most heterogeneous
category of plastids and are classified entirely on the
structure of the carotenoid-bearing components present
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 27
plastoglobule
tonoplast
chloroplast envelope
plasma
membrane
tonoplast
vacuole
stroma thylakoid
granum
peripheral reticulum
n
cell
wall
stroma
cell wall
m
granum
A
0.55 mm
B
1 mm
FIGURE 2.9
A, chloroplasts along the cell wall in a leaf cell of sheperd’s purse (Capsella bursa-pastoris). Mitochondria (m) are
closely associated spatially with the chloroplasts. B, chloroplast with grana seen in profile. From a leaf of tobacco
(Nicotiana tabacum). (B, from Esau, 1977.)
in the mature plastid (Sitte et al., 1980). Most belong to
one of four types: (1) globular chromoplasts, with
many carotenoid-bearing plastoglobuli (Fig. 2.11A).
Remnants of thylakoids also may be present. The plastoglobuli often are concentrated in the peripheral
stroma beneath the envelope (petals of Ranunculus
repens and yellow fruits of Capsicum, perianth of
Tulipa, Citrus fruit); (2) membranous chromoplasts,
which are characterized by a set of up to 20 concentric
(double) carotenoid-containing membranes (Fig. 2.11B)
(Narcissus and Citrus sinensis petals); (3) tubular
chromoplasts, in which the carotenoids are incorporated into filamentous lipoprotein “tubules” (Fig. 2.11C)
(red fruits of Capsicum, rose hypanthium; Tropaeolum
petals; Knoth et al., 1986); (4) crystalline chromoplasts, which contain crystalline inclusions of pure
carotene (Fig. 2.11D) (β-carotene in Daucus, carrot,
roots and lycopene in Solanum lycopersicum, tomato,
fruit). Carotene crystals, commonly called pigment
bodies, originate within thylakoids and remain
28 | Esau’s Plant Anatomy, Third Edition
A
5 mm
B
FIGURE 2.10
Chloroplast structure. A, with the light microscope grana within chloroplasts appear as dots. These chloroplasts are
from a cotyledon of Solanum lycopersicum. B, an electron micrograph of a chloroplast from a bundle-sheath cell of
a Zea leaf showing grana from the surface. (A, from Hagemann, 1960.)
surrounded by the plastid envelope during all stages
of development. Globular chromoplasts are the most
common type and are regarded as the oldest and most
primitive in evolutionary terms (Camara et al., 1995).
Chromoplasts may develop from previously existing
green chloroplasts by a transformation in which the
chlorophyll and thylakoid membranes of the chloroplast
disappear and masses of carotenoids accumulate, as
occurs during the ripening of many fruits (Ziegler et al.,
1983; Kuntz et al., 1989; Marano and Carrillo, 1991,
1992; Cheung et al., 1993; Ljubešić et al., 1996). Interestingly these changes apparently are accompanied by the
disappearance of plastid ribosomes and rRNAs but
not of the plastid DNA, which remains unchanged
(Hansmann et al., 1987; Camara et al., 1989; Marano and
Carrillo, 1991). With the loss of plastid ribosomes
and rRNAs, protein synthesis can no longer occur in
the chromoplast, indicating that it is necessary for
chromoplast-specific proteins to be coded for in the
nucleus and then imported into the developing chromoplast. Chromoplast development is not an irreversible
phenomenon, however. For example, the chromoplasts
of citrus fruit (Goldschmidt, 1988) and of the carrot root
(Grönegress, 1971) are capable of reverse differentiation
into chloroplasts; they lose the carotene pigment and
develop a thylakoid system, chlorophyll, and photosynthetic apparatus.
The precise functions of chromoplasts are not well
understood, although at times they act as attractants to
insects and other animals with which they coevolved,
playing an essential role in the cross-pollination of flow-
ering plants and the dispersal of fruit and seeds (Raven
et al., 2005).
Leucoplasts Are Nonpigmented Plastids
Structurally the least differentiated of mature plastids,
leucoplasts (Fig. 2.12) generally have a uniform granular stroma, several nucleoids, and, despite reports to the
contrary, typical 70S ribosomes. They lack an elaborate
system of inner membranes (Carde, 1984; Miernyk,
1989). Some store starch (amyloplasts; Fig. 2.13),
others store proteins (proteinoplasts), fats (elaioplasts), or combinations of these products. Amyloplasts
are classified as simple or compound (Shannon, 1989).
Simple amyloplasts, such as those of the potato tuber,
contain a single starch grain, whereas compound amyloplasts contain several often tightly packed starch
grains as in the endosperm of oats and rice. The starch
grains of the potato tuber may become so large that the
envelope is ruptured (Kirk and Tilney-Bassett, 1978).
The compound amyloplasts in rootcaps play an essential
role in gravity perception (Sack and Kiss, 1989; Sack,
1997).
All Plastids Are Derived Initially from Proplastids
Proplastids are small, colorless plastids found in undifferentiated regions of the plant body such as root and
shoot apical meristems (Mullet, 1988). Zygotes contain
proplastids that are the ultimate precursors of all plastids within an adult plant. In most angiosperms the
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 29
A
B
0.5 µm
0.5 µm
ob
cr
D
C
0.5 µm
0.5 µm
FIGURE 2.11
Types of chromoplasts. A, globular chromoplasts from Tagetes (marigold) petal; B, membranous chromoplast of
Narcissus pseudonarcissus flower; C, tubular chromoplast of Palisota barteri fruit; D, crystalline chromoplast of
Solanum lycopersicum fruit. Details: cr, crystalloids; ob, oil body. (B, reprinted from Hansmann et al., 1987. © 1987,
with permission from Elsevier.; C, from Knoth et al., 1986, Fig. 7. © 1986 Springer-Verlag; D, from Mohr, 1979, by
permission of Oxford Unïversity Press.)
proplastids of the zygote come exclusively from the
cytoplasm of the egg cell (Nakamura et al., 1992). In
conifers, however, the proplastids of the zygote are
derived from those carried by the sperm cell. In either
case the consequence is that the plastid genome of an
individual plant typically is inherited from a single
parent. Since all the plastids in an adult plant are derived
from a single parent, all plastids (whether chloroplasts,
chromoplasts, or leucoplasts) within an individual
plant have identical genomes (dePamphilis and Palmer,
1989). Each proplastid contains a single circular DNA
molecule.
30 | Esau’s Plant Anatomy, Third Edition
FIGURE 2.12
Leucoplasts clustered around the nucleus in epidermal
cells of a Zebrina leaf. (×620.)
0.83 mm
FIGURE 2.14
Dividing chloroplast in Beta vulgaris leaf. Had the division process continued, the two daughter plastids would
have separated at the narrow constriction, or isthmus.
Three peroxisomes can be seen to the right of the
constriction.
1.3 mm
FIGURE 2.13
Amyloplast, a type of leucoplast, from the embryo sac
of soybean (Glycine max . The round, clear bodies are
starch grains. The smaller, dense bodies are oil bodies.
Amyloplasts are involved with the synthesis and longterm storage of starch in seeds and storage organs, such
as potato tubers. (Courtesy of Roland R. Dute.)
As mentioned previously, plastids reproduce by
binary fission, the process of dividing into equal halves,
which is characteristic of bacteria (Oross and
Possingham, 1989). In meristematic cells the division of
proplastids roughly keeps pace with cell division. The
proplastids must divide before the cells divide. The
plastid population of mature cells typically exceeds that
of the original proplastid population. The greater proportion of the fi nal plastid population may be derived
from the division of mature plastids during the period
of cell expansion. Although plastid division apparently
is controlled by the nucleus (Possingham and Lawrence,
1983), a close interaction exists between plastid DNA
replication and plastid division.
Plastid division is initiated by a constriction in the
middle of the plastid (Fig. 2.14). With continued narrowing of the constriction, the two daughter plastids come
to be joined by a narrow isthmus, which eventually
breaks. The envelope membranes of the daughter plastids then reseal. The constriction process is caused by
contractile rings referred to as plastid-dividing rings,
which are discernible with the electron microscope as
electron-dense bands. There are two concentric plastiddividing rings, an outer ring on the cytosolic face of the
plastid outer membrane and an inner ring on the stromal
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 31
Amyloplast
G
Chromoplast
F
0.84 mm
Proplastid
A
B
prolamellar
body
thylakoid
envelope
Mature
chloroplast
E
FIGURE 2.15
C
Etiolated chloroplast with a prolamellar body in a leaf
cell of sugarcane (Saccharum officinarum). Ribosomes
are conspicuous in the plastid. (Courtesy of W. M.
Laetsch.)
D
FIGURE 2.16
face of the plastid inner membrane. Prior to the appearance of the plastid-dividing rings, two cytoskeletal-like
proteins, FtsZ1 and FtsZ2—homologs of the bacterial cell
division FtsZ protein—assemble into a ring at the future
division site in the stroma within the plastid envelope.
It has been suggested that the FtsZ ring determines the
division region (Kuroiwa et al., 2002). Molecular analysis
of chloroplast division indicates that the mechanism of
plastid division has evolved from bacterial cell division (Osteryoung and Pyke, 1998; Osteryoung and
McAndrew, 2001; Miyagishima et al., 2001).
If the development of a proplastid into a more highly
differentiated form is arrested by the absence of light,
it may form one or more prolamellar bodies (Fig.
2.15), which are quasi-crystalline bodies composed of
tubular membranes (Gunning, 2001). Plastids containing prolamellar bodies are called etioplasts (Kirk and
Tilney-Bassett, 1978). Etioplasts form in leaf cells of
plants grown in the dark. During subsequent development of etioplasts into chloroplasts in the light, the
membranes of the prolamellar bodies develop into thylakoids. Carotenoid synthesis has been demonstrated to
be required for the formation of prolamellar bodies in
etiolated seedlings of Arabidopsis (Park et al., 2002). In
nature, the proplastids in the embryos of some seeds
first develop into etioplasts; then, upon exposure to
light, the etioplasts develop into chloroplasts. The
various kinds of plastids are remarkable for the relative
ease with which they can change from one type to
another (Fig. 2.16).
Plastid developmental cycle, beginning with the development of a chloroplast from a proplastid (A). Initially
the proplastid contains few or no internal membranes.
B–D, as the proplastid differentiates, flattened vesicles
develop from the inner membrane of the plastid envelope and eventually align themselves into grana and
stroma thylakoids. E, the thylakoid system of the mature
chloroplast appears discontinuous with the envelope.
F, G, proplastids may also develop into chromoplasts
and leucoplasts. The leucoplast shown here is a starchsynthesizing amyloplast. Note that chromoplasts may be
formed from proplastids, chloroplasts, or leucoplasts.
The various kinds of plastids can change from one type
to another (dashed arrows). (From Raven et al., 2005.)
❙ MITOCHONDRIA
Mitochondria, like plastids, are bounded by two membranes (Figs. 2.17 and 2.18). The inner membrane is
convoluted inwardly into numerous folds known as
cristae (singular: crista), which greatly increase the
surface area available to enzymes and the reactions associated with them. Mitochondria are generally smaller
than plastids, measuring about half a micrometer in
diameter and exhibiting great variation in length and
shape.
Mitochondria are the sites of respiration, a process
involving the release of energy from organic molecules
and its conversion to molecules of ATP (adenosine
32 | Esau’s Plant Anatomy, Third Edition
Outer membrane
Inner membrane
Intermembrane space
Crista
Matrix
FIGURE 2.17
Three-dimensional structure of a mitochondrion. The
inner of the two membranes bounding the mitochondrion fold inward, forming the cristae. Many of the
enzymes and electron carriers involved in respiration
are present in the cristae. (From Raven et al., 2005.)
crista
0.14 mm
B
cw
cristae
envelope
cw
A
0.70 mm
FIGURE 2.18
Mitochondria. A, in a leaf cell of tobacco (Nicotiana
tabacum). The envelope consists of two membranes,
and the cristae are embedded in a dense stroma. B,
mitochondrion in a leaf cell of spinach (Spinacia oleracea), in a section revealing some strands of DNA in the
nucleoid. Detail: cw, cell wall.
triphosphate), the principal immediate energy source
for the cell (Mackenzie and McIntosh, 1999; Møller,
2001; Bowsher and Tobin, 2001). Within the innermost
compartment, surrounding the cristae, is the matrix, a
dense solution containing enzymes, coenzymes, water,
phosphate, and other molecules involved with respiration. Whereas the outer membrane is fairly permeable
to most small molecules, the inner one is relatively
impermeable, permitting the passage of only certain
molecules, such as pyruvate and ATP, while preventing
the passage of others. Some enzymes of the citric-acid
cycle are found in solution in the matrix. Other citric
acid-cycle enzymes and the components of the electrontransport chain are built into the surfaces of the cristae.
Most plant cells contain hundreds of mitochondria, the
number of mitochondria per cell being related to the
cell’s demand for ATP.
Mitochondria are in constant motion and appear to
move freely in the streaming cytoplasm from one part
of the cell to another; they also fuse and divide by binary
fission (Arimura et al., 2004), involving dividing rings
reminiscent of the plastid-dividing rings (Osteryoung,
2000). Movement of mitochondria in cultured cells of
tobacco (Nicotiana tabacum) has been shown to
involve an actin-myosin-based system (Van Gestel et al.,
2002). Mitochondria tend to congregate where energy
is required. In cells in which the plasma membrane is
very active in transporting materials into or out of the
cell, the mitochondria often can be found arrayed along
the membrane surface.
Mitochondria, like plastids, are semiautonomous
organelles, containing the components necessary for
the synthesis of some of their own proteins. One or
more DNA-containing nucleoids and many 70S ribosomes similar to those of bacteria are found in the
matrix (Fig. 2.18). The DNA is not associated with histones. Thus in plant cells genetic information is found
in three different compartments: nucleus, plastid, and
mitochondrion. The mitochondrial genomes of plants
are much larger (200–2400 kb) than those of animals
(14–42 kb), fungi (18–176 kb), and plastids (120–200 kb)
(Backert et al., 1997; Giegé and Brennicke, 2001). Their
structural organization is not fully understood. Linear
and circular DNA molecules of variable size as well as
more complex DNA molecules are consistently present
(Backert et al., 1997).
Mitochondria are widely accepted to have evolved
from free-living α-proteobacteria through the process
of endosymbiosis (Gray, 1989). As with the chloroplast,
in the course of evolution the DNA of the mitochondria
was massively transferred to the nucleus (Adams et al.,
2000; Gray, 2000). Evidence also indicates that
some genetic information has been transferred from
chloroplasts to mitochondria over long periods of evolutionary time (Nugent and Palmer, 1988; Jukes and
Osawa, 1990; Nakazono and Hirai, 1993) and possibly
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 33
from the nucleus to the mitochondria (Schuster and
Brennicke, 1987; Marienfeld et al., 1999). Only about 30
proteins are encoded in plant mitochondrial genomes.
By contrast, about 4000 proteins encoded in the nucleus
are estimated to be imported from the cytosol. Nuclear
encoded mitochondrial proteins contain signal peptides
called presequences at their N-terminus to direct them
into the mitochondria (Braun and Schmitz, 1999; Mackenzie and McIntosh, 1999; Giegé and Brennicke,
2001).
Genetic information found only in mitochondrial
DNA may have an effect on cell development. Most
notable is cytoplasmic male sterility, a maternally inherited (mitochondrial DNA is maternally inherited) trait
that prevents the production of functional pollen but
does not affect female fertility (Leaver and Gray, 1982).
Because it prevents self-pollination, the cytoplasmic
male sterility phenotype has been widely used in the
commercial production of F1 hybrid seed (e.g., in maize,
onions, carrots, beets, and petunias).
Mitochondria have come to be regarded as key players in the regulation of programmed cell death, called
apoptosis, in animal cells (Chapter 5; Desagher and
Martinou, 2000; Ferri and Kroemer, 2001; Finkel, 2001).
A primary cellular trigger for apoptosis is the release of
cytochrome c from the mitochondrial intermembrane
space. Release of the cytochrome c appears to be a
critical event for the activation of catabolic proteases
called caspases (apoptosis-specific cysteine proteases).
Although mitochondria may play a role in plant programmed cell death, it is unlikely that released cytochrome c is involved in that process (Jones, 2000; Xu
and Hanson, 2000; Young and Gallie, 2000; Yu et al.,
2002; Balk et al., 2003; Yao et al., 2004).
❙ PEROXISOMES
Unlike plastids and mitochondria, which are bounded
by two membranes, peroxisomes (also called microbodies) are spherical organelles bounded by a single
membrane (Figs. 2.14 and 2.19; Frederick et al., 1975;
Olsen, 1998). Peroxisomes differ most notably from
plastids and mitochondria, however, in their lack of
DNA and ribosomes. Consequently all peroxisomal proteins are nuclear-encoded, and at least the matrix proteins are synthesized on ribosomes in the cytosol and
then transported into the peroxisome. A subset of peroxisomal membrane proteins might be targeted first to
the endoplasmic reticulum, and from there to the organelle by vesicle-mediated transport (Johnson and Olsen,
2001). Peroxisomes range in size from 0.5 to 1.5 μm.
They lack internal membranes and have a granular interior, which sometimes contains an amorphous or crystalline body composed of protein. According to the
prevailing view, peroxisomes are self-replicating organ-
elles, new peroxisomes arising from preexisting ones
by division. The existence of a vesicle-mediated pathway
from the endoplasmic reticulum to the peroxisomes has
led some workers to speculate that these organelles may
also be generated de novo (Kunau and Erdmann, 1998;
Titorenko and Rachubinski, 1998; Mullen et al., 2001),
a view strongly challenged by others (Purdue and
Lazarow, 2001). Biochemically peroxisomes are characterized by the presence of at least one hydrogen
peroxide–producing oxidase and catalase for the
removal of the hydrogen peroxide (Tolbert, 1980; Olsen,
1998). As noted by Corpas et al. (2001), an important
property of peroxisomes is their “metabolic plasticity,”
in that their enzymatic content can vary, depending on
the organism, cell type or tissue type, and environmental conditions. Peroxisomes perform a wide array of
metabolic functions (Hu et al., 2002).
Two very different types of peroxisome have been
studied extensively in plants (Tolbert and Essner, 1981;
Trelease, 1984; Kindl, 1992). One of them occurs in
green leaves, where it plays an important role in glycolic
acid metabolism, which is associated with photorespiration, a process that consumes oxygen and releases
carbon dioxide. Photorespiration involves cooperative
interaction among peroxisomes, mitochondria, and
chloroplasts; hence these three organelles commonly
are closely associated spatially with one another (Fig.
2.19A). The biological function of photorespiration
remains to be determined (Taiz and Zeiger, 2002).
The second type of peroxisome is found in the endosperm or cotyledons of germinating seeds, where it
plays an essential role in the conversion of fats to carbohydrates by a series of reactions known as the glyoxylate cycle. Appropriately these peroxisomes are also
called glyoxysomes. The two types of peroxisome are
interconvertible (Kindl, 1992; Nishimura et al., 1993,
1998). For example, during the early stages of germination the cotyledons of some seeds are essentially
deprived of light. As the cotyledons gradually become
exposed to light, they may become green. With the
depletion of fat and the appearance of chloroplasts, the
glyoxysomes are converted to leaf-type peroxisomes.
Glyoxysomal properties may reappear as the tissues
undergo senescence.
Several studies have revealed that plant peroxisomes,
like plastids and mitochondria, are motile organelles
whose movement is actin dependent (Collings et al.,
2002; Jedd and Chua, 2002; Mano et al., 2002; Mathur
et al., 2002). The peroxisomes in leek (Allium porrum)
and Arabidopsis have been shown to undergo dynamic
movements along bundles of actin fi laments (Collings
et al., 2002; Mano et al., 2002), those in Arabidopsis
reaching peak velocities approaching 10 μm·s−1 (Jedd
and Chua, 2002). Moreover the peroxisomes in Arabidopsis have been shown to be driven by myosin motors
(Jedd and Chua, 2002).
34 | Esau’s Plant Anatomy, Third Edition
1 mm
starch
A
chloroplast
peroxisome
mitochondrion
ribosome
crista
B
0.5 mm
m
FIGURE 2.19
Organelles in leaf cells of sugar beet (Beta vulgaris, A) and tobacco (Nicotiana tabacum, B). The unit membranes
enclosing the peroxisomes may be contrasted with the double-membraned envelopes of the other organelles. The
peroxisome in B contains a crystal. Some ribosomes are perceptible in the chloroplast in A and in the mitochondrion
in B. (From Esau, 1977.)
❙ VACUOLES
Together with the presence of plastids and a cell wall,
the vacuole is one of the three characteristics that distinguish plant cells from animal cells. As mentioned
previously, vacuoles are organelles bounded by a single
membrane, the tonoplast, or vacuolar membrane
(Fig. 2.2). They are multifunctional organelles and are
widely diverse in form, size, content, and functional
dynamics (Wink, 1993; Marty, 1999). A single cell may
contain more than one kind of vacuole. Some vacuoles
function primarily as storage organelles, others as lytic
compartments. The two types of vacuole can be characterized by the presence of specific tonoplast integral
(intrinsic) proteins (TIPs): for example, whereas α-TIP
is associated with the tonoplasts of protein-storage vacu-
oles, γ-TIP localizes to the tonoplasts of lytic vacuoles.
Both types of TIP may colocalize to the same tonoplast
of large vacuoles, apparently the result of merger of the
two types of vacuole during cell enlargement (Paris
et al., 1996; Miller and Anderson, 1999).
Many meristematic plant cells contain numerous
small vacuoles. As the cell enlarges, the vacuoles increase
in size and fuse into a single large vacuole (Fig. 2.20).
Most of the increase in size of the cell in fact involves
enlargement of the vacuoles. In the mature cell as much
as 90% of the volume may be taken up by the vacuole,
with the rest of the cytoplasm consisting of a thin
peripheral layer closely pressed against the cell wall.
By filling such a large proportion of the cell with “inexpensive” (in terms of energy) vacuolar contents, plants
not only save “expensive” nitrogen-rich cytoplasmic
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 35
vacuole
vacuole
0.56 mm
FIGURE 2.21
1.33 mm
Tannin-containing vacuole in leaf cell of the sensitive
plant (Mimosa pudica). The electron-dense tannin literally fi lls the central vacuole of this cell.
FIGURE 2.20
Parenchyma cell from a tobacco (Nicotiana tabacum)
leaf, with its nucleus “suspended” in the middle of the
vacuole by dense strands of cytoplasm. The dense granular substance in the nucleus is chromatin.
material but also acquire a large surface area between
the thin layer of nitrogen-rich cytoplasm and the
protoplast’s external environment (Wiebe, 1978). Being
a selectively permeable membrane, the tonoplast is
involved with the regulation of osmotic phenomena
associated with the vacuoles. A direct consequence of
this strategy is the development of tissue rigidity, one
of the principal roles of the vacuole and tonoplast.
The principal component of the non–protein-storing
vacuoles is water, with other components varying
according to the type of plant, organ, and cell and their
developmental and physiological state (Nakamura and
Matsuoka, 1993; Wink, 1993). In addition to inorganic
ions such as Ca2+ , Cl−, K + , Na + , NO3−, and PO42−, such
vacuoles commonly contain sugars, organic acids, and
amino acids, and the aqueous solution commonly is
called cell sap. Sometimes the concentration of a particular substance in the vacuole is sufficiently great for
it to form crystals. Calcium oxalate crystals, which can
assume different forms (Chapter 3), are especially
common. In most cases vacuoles do not synthesize the
molecules that they accumulate but must receive them
from other parts of the cytoplasm. The transport of
metabolites and inorganic ions across the tonoplast is
strictly controlled to ensure optimal functioning of the
cell (Martinoia, 1992; Nakamura and Matsuoka, 1993;
Wink, 1993).
Vacuoles are important storage compartments for
various metabolites. Primary metabolites—substances
that play a basic role in cell metabolism—such as sugars
and organic acids are stored only temporarily in the
vacuole. In photosynthesizing leaves of many species,
for example, much of the sugar produced during the day
is stored in the mesophyll cell vacuoles and then moved
out of the vacuoles during the night for export to other
parts of the plant. In CAM plants, malic acid is stored
in the vacuoles during the night and released from the
vacuoles and decarboxylated during the day, the CO2
then becoming assimilated by the Calvin cycle in the
chloroplasts (Kluge et al., 1982; Smith, 1987). In seeds,
vacuoles are a primary site for the storage of proteins
(Herman and Larkins, 1999).
Vacuoles also sequester toxic secondary metabolites,
such as nicotine, an alkaloid, and tannins, phenolic compounds, from the rest of the cytoplasm (Fig. 2.21). Secondary metabolites play no apparent role in the plant’s
36 | Esau’s Plant Anatomy, Third Edition
primary metabolism. Such substances may be sequestered permanently in the vacuoles. A great many of the
secondary metabolites accumulated in the vacuoles are
toxic not only to the plant itself but also to pathogens,
parasites, and/or herbivores, and therefore they play an
important role in plant defense. Some of the secondary
metabolites stored in the vacuoles are nontoxic but are
converted upon hydrolysis to such highly toxic derivatives as cyanide, mustard oils, and aglycones when the
vacuoles are ruptured (Matile, 1982; Boller and Wiemken,
1986). Thus detoxification of the cytoplasm and the
storage of defensive chemicals may be regarded as additional functions of vacuoles.
The vacuole is often the site of pigment deposition.
The blue, violet, purple, dark red, and scarlet colors of
plant cells are usually caused by a group of pigments
known as the anthocyanins. These pigments frequently are confi ned to epidermal cells. Unlike most
other plant pigments (e.g., chlorophylls, carotenoids),
the anthocyanins are readily soluble in water and are
found in solution in the vacuole. They are responsible
for the red and blue colors of many fruits (grapes, plums,
cherries) and vegetables (radishes, turnips, cabbages),
and a host of flowers (geraniums, delphiniums, roses,
petunias, peonies), and presumably serve to attract
animals for pollination and seed dispersal. Anthocyanin
has been implicated with the sequestration of molybdenum in vacuoles of peripheral cell layers of Brassica
seedlings (Hale et al., 2001). In a restricted number of
plant families, another class of water-soluble pigments,
the nitrogen-containing betalains, is responsible for
some of the yellow and red colors. These plants, all
members of the order Chenopodiales, lack anthocyanins. The red color of beets and Bougainvillea flowers
is due to the presence of betacyanins (red betalains).
The yellow betalains are called betaxanthins (Piattelli,
1981).
Anthocyanins are also responsible for the brilliant
red colors of some leaves in autumn. These pigments
form in response to cold, sunny weather, when leaves
stop producing chlorophyll. As the chlorophyll that is
present disintegrates, the newly formed anthocyanins
are unmasked. In leaves that do not form anthocyanin
pigments, the breakdown of chlorophyll in autumn may
unmask the more stable yellow-to-orange carotenoid
pigments already present it the chloroplasts. The most
spectacular autumnal coloration develops in years when
cool, clear weather prevails in the fall (Kozlowski and
Pallardy, 1997).
What role is played by anthocyanins found in leaves?
In red-osier dogwood (Cornus stolonifera), anthocyanins form a pigment layer in the palisade mesophyll
layer in autumn, decreasing light capture by the chloroplasts prior to leaf fall. It has been suggested that this
optical masking of chlorophyll by the anthocyanins
reduces the risk of photo-oxidative damage to the leaf
cells as they senesce, damage that otherwise might
lower the efficiency of nutrient retrieval from the senescing leaves (Feild et al., 2001). In addition to protecting
leaves from photo-oxidative damage, evidence indicates that anthocyanins protect against photoinhibition
(Havaux and Kloppstech, 2001; Lee, D. W., and Gould,
2002; Steyn et al., 2002), a decline in photosynthetic
efficiency resulting from excess excitation arriving at
the reaction center of photosystem II. Photoinhibition
is common in understory plants and occurs when they
are suddenly exposed to patches of full sunlight (sunflecks) that pass through momentary openings in the
upper canopy as the leaves flutter in the breeze (Pearcy,
1990).
As lytic compartments, vacuoles are involved with
the breakdown of macromolecules and the recycling of
components within the cell. Entire organelles, such as
senescent plastids and mitochondria, may be engulfed
and subsequently degraded by vacuoles containing large
numbers of hydrolytic and oxidizing enzymes. The large
central vacuole can sequester hydrolases, which upon
breakdown of the tonoplast can result in complete autolysis of the cytoplasm, as during programmed cell death
of differentiating tracheary elements (Chapter 10).
Because of this digestive activity the so-called lytic vacuoles are comparable in function with the organelles
known as lysosomes in animal cells.
New vacuoles have long been considered to arise
from dilation of specialized regions of the smooth ER or
from vesicles derived from the Golgi apparatus. Most
evidence supports de novo formation of vacuoles from
the ER (Robinson, 1985; Hörtensteiner et al., 1992;
Herman et al., 1994).
❙ RIBOSOMES
Ribosomes are small particles, only about 17 to 23
nanometers in diameter (Fig. 2.22), consisting of protein
and RNA (Davies and Larkins, 1980). Although the
number of protein molecules in ribosomes greatly
exceeds the number of RNA molecules, RNA constitutes
about 60% of the mass of a ribosome. They are the sites
at which amino acids are linked together to form proteins and are abundant in the cytoplasm of metabolically active cells (Lake, 1981). Each ribosome consists
of two subunits, one small and the other large, composed of specific ribosomal RNA and protein molecules.
Ribosomes occur both freely in the cytosol and attached
to the endoplasmic reticulum and outer surface of the
nuclear envelope. They are by far the most numerous of
cellular structures and are also found in nuclei, plastids,
and mitochondria. As mentioned previously, the ribosomes of plastids and mitochondria are similar in size
to those of bacteria.
Ribosomes actively involved in protein synthesis
occur in clusters or aggregates called polysomes, or
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 37
mitochondria, plastids, or peroxisomes remain free in
the cytosol. The polypeptides released from the free
polysomes either remain in the cytosol or are targeted
to the appropriate cellular component by a targeting
sequence (Holtzman, 1992). Membrane-bound and free
ribosomes are both structurally and functionally identical, differing from one another only in the proteins they
are making at any given time.
REFERENCES
ADAMS , K. L., D. O. DALEY, Y.-L. QIU, J. WHELAN, and J. D.
PALMER . 2000. Repeated, recent and diverse transfers of a
mitochondrial gene to the nucleus in flowering plants. Nature
408, 354–357.
ARIMURA , S.-I., A. HIRAI, and N. TSUTSUMI. 2001. Numerous and
highly developed tubular projections from plastids observed
in tobacco epidermal cells. Plant Sci. 160, 449–454.
200 nm
A
100 nm
B
FIGURE 2.22
Ribosomes. A, in bundle-sheath cell of a maize (Zea
mays) leaf. Arrow points to a bundle of actin fi laments.
B, a polysome (polyribosome) attached to the surface
of endoplasmic reticulum in a tobacco (Nicotiana
tabacum) leaf cell. (B, from Esau, 1977.)
polyribosomes (Fig. 2.22), united by the messenger
RNA molecules carrying the genetic information from
the nucleus. The amino acids from which the proteins
are synthesized are brought to the polysomes by transfer
RNAs located in the cytosol. The synthesis of protein,
known as translation, consumes more energy than any
other biosynthetic process. That energy is provided by
the hydrolysis of guanosine triphosphate (GTP).
The synthesis of polypeptides (proteins) encoded by
nuclear genes is initiated on polysomes located in the
cytosol and follows one of two divergent pathways. (1)
Those polysomes involved in the synthesis of polypeptides destined for the endoplasmic reticulum become
associated with the endoplasmic reticulum early in the
translational process. The polypeptides and their associated polysomes are directed to the endoplasmic reticulum by a targeting signal, or signal peptide, located at
the amino end of each polypeptide. The polypeptides
are transferred across the membrane into the lumen of
the ER (or are inserted into it, in the case of integral
proteins) as polypeptide synthesis continues. (2) Those
polysomes involved with the synthesis of polypeptides
destined for the cytosol or for import into the nucleus,
ARIMURA , S.-I., J. YAMAMOTO, G. P. AIDA , M. NAKAZONO, and
N. TSUTSUMI. 2004. Frequent fusion and fission of plant mitochondria with unequal nucleoid distribution. Proc. Natl. Acad.
Sci. USA 101, 7805–7808.
B ACKERT, S., B. L. NIELSEN, and T. BÖRNER . 1997. The mystery of
the rings: Structure and replication of mitochondrial genomes
from higher plants. Trends Plant Sci. 2, 477–483.
B ALK, J., S. K. CHEW, C. J. LEAVER , and P. F. MCCABE. 2003. The
intermembrane space of plant mitochondria contains a DNase
activity that may be involved in programmed cell death. Plant
J. 34, 573–583.
B ASKIN, T. I. 2000. The cytoskeleton. In: Biochemistry and
Molecular Biology of Plants, pp. 202–258, B. B. Buchanan, W.
Gruissem, and R. L. Jones, eds. American Society of Plant
Physiologists, Rockville, MD.
B ATTEY, N. H., N. C. JAMES , A. J. GREENLAND, and C. BROWNLEE.
1999. Exocytosis and endocytosis. Plant Cell 11, 643–659.
BOLLER , T., and A. WIEMKEN. 1986. Dynamics of vacuolar compartmentation. Annu. Rev. Plant Physiol. 37, 137–164.
BÖLTER , B., and J. SOLL . 2001. Ion channels in the outer membranes of chloroplasts and mitochondria: Open doors or
regulated gates? EMBO J. 20, 935–940.
BONIOTTI, M. B., and M. E. GRIFFITH. 2002. “Cross-talk” between
cell division cycle and development in plants. Plant Cell 14,
11–16.
BOWSHER , C. G., and A. K. TOBIN. 2001. Compartmentation of
metabolism within mitochondria and plastids. J. Exp. Bot. 52,
513–527.
BRAUN, H.-P., and U. K. SCHMITZ. 1999. The protein-import
apparatus of plant mitochondria. Planta 209, 267–274.
BRUCE, B. D. 2000. Chloroplast transit peptides: Structure, function and evolution. Trends Cell Biol. 10, 440–447.
CAMARA , B., J. BOUSQUET, C. CHENICLET, J.-P. CARDE, M. KUNTZ,
J.-L. EVRARD, and J.-H. WEIL . 1989. Enzymology of isoprenoid
38 | Esau’s Plant Anatomy, Third Edition
biosynthesis and expression of plastid and nuclear genes
during chromoplast differentiation in pepper fruits (Capsicum
annuum). In: Physiology, Biochemistry, and Genetics of Nongreen
Plastids, pp. 141–156, C. D. Boyer, J. C. Shannon, and R. C.
Hardison, eds. American Society of Plant Physiologists,
Rockville, MD.
CAMARA , B., P. HUGUENEY, F. BOUVIER , M. KUNTZ, and R.
MONÉGER . 1995. Biochemistry and molecular biology of chromoplast development. Int. Rev. Cytol. 163, 175–247.
CARDE, J.-P. 1984. Leucoplasts: A distinct kind of organelles
lacking typical 70S ribosomes and free thylakoids. Eur. J. Cell
Biol. 34, 18–26.
CHEUNG, A. Y., T. MCNELLIS , and B. PIEKOS. 1993. Maintenance
of chloroplast components during chromoplast differentiation in the tomato mutant Green Flesh. Plant Physiol. 101,
1223–1229.
CHRISPEELS , M. J., N. M. CRAWFORD, and J. I. SCHROEDER. 1999.
Proteins for transport of water and mineral nutrients across
the membranes of plant cells. Plant Cell 11, 661–676.
CLINE, K., R. HENRY, C.-J. LI, and J.-G. YUAN. 1993. Multiple
pathways for protein transport into or across the thylakoid
membrane. EMBO J. 12, 4105–4114.
uptake of organic solutes and stress resistance. Plant Sci. 161,
391–404.
BOER , B. G. W., and J. A. H. MURRAY. 2000. Triggering the
cell cycle in plants. Trends Cell Biol. 10, 245–250.
DEN
DE PAMPHILIS ,
C. W., and J. D. PALMER . 1989. Evolution and function of plastid DNA: A review with special reference to
nonphotosynthetic plants. In: Physiology, Biochemistry, and
Genetics of Nongreen Plastids, pp. 182–202, C. D. Boyer, J. C.
Shannon, and R. C. Hardison, eds. American Society of Plant
Physiologists, Rockville, MD.
DESAGHER , S., and J.-C. MARTINOU. 2000. Mitochondria as the
central control point of apoptosis. Trends Cell Biol. 10,
369–377.
DINGWALL , C., and R. L ASKEY. 1992. The nuclear membrane.
Science 258, 942–947.
ESAU, K. 1977. Anatomy of Seed Plants, 2nd ed. Wiley, New
York.
FEILD, T. S., D. W. LEE, and N. M. HOLBROOK. 2001. Why leaves
turn red in autumn. The role of anthocyanins in senescing
leaves of red-osier dogwood. Plant Physiol. 127, 566–574.
FERRI, K. F., and G. KROEMER . 2001. Mitochondria—The suicide
organelles. BioEssays 23, 111–115.
CLOWES , F. A. L., and B. E. JUNIPER . 1968. Plant Cells. Blackwell
Scientific, Oxford.
FINKEL , E. 2001. The mitochondrion: Is it central to apoptosis?
Science 292, 624–626.
COLLINGS , D. A., J. D. I. HARPER , J. MARC , R. L. OVERALL , and
R. T. MULLEN. 2002. Life in the fast lane: Actin-based motility
of plant peroxisomes. Can. J. Bot. 80, 430–441.
FLÜGGE, U.-I. 1990. Import of proteins into chloroplasts. J. Cell
Sci. 96, 351–354.
COOKE, T. J., and B. LU. 1992. The independence of cell shape
and overall form in multicellular algae and land plants: Cells
do not act as building blocks for constructing plant organs.
Int. J. Plant Sci. 153, S7–S27.
CORPAS , F. J., J. B. B ARROSO, and L. A. DEL RÍO. 2001. Peroxisomes as a source of reactive oxygen species and nitric oxide
signal molecules in plant cells. Trends Plant Sci. 6, 145–150.
CUNNINGHAM, F. X., JR ., and E. GANTT. 1998. Genes and enzymes
of carotenoid biosynthesis in plants. Annu. Rev. Plant Physiol.
Plant Mol. Biol. 49, 557–583.
CUTLER , S., and D. EHRHARDT. 2000. Dead cells don’t dance:
Insights from live-cell imaging in plants. Curr. Opin. Plant Biol.
3, 532–537.
FRANKLIN, A. E., and W. Z. CANDE. 1999. Nuclear organization
and chromosome segregation. Plant Cell 11, 523–534.
FREDERICK, S. E., P. J. GRUBER , and E. H. NEWCOMB . 1975. Plant
microbodies. Protoplasma 84, 1–29.
FRICKER , M. D., and K. J. OPARKA . 1999. Imaging techniques in
plant transport: Meeting review. J. Exp. Bot. 50(suppl. 1), 1089–
1100.
FULGOSI, H., and J. SOLL . 2001. A gateway to chloroplasts—
Protein translocation and beyond. J. Plant Physiol. 158,
273–284.
GAIDAROV, I., F. SANTINI, R. A. WARREN, and J. H. KEEN. 1999.
Spatial control of coated-pit dynamics in living cells. Nature
Cell Biol. 1, 1–7.
GANT, T. M., and K. L. WILSON. 1997. Nuclear assembly. Annu.
Rev. Cell Dev. Biol. 13, 669–695.
D’AMATO, F. 1998. Chromosome endoreduplication in plant
tissue development and function. In: Plant Cell Proliferation and
Its Regulation in Growth and Development, pp. 153–166, J. A.
Bryant and D. Chiatante, eds. Wiley, Chichester.
GERACE, L., and R. FOISNER . 1994. Integral membrane proteins
and dynamic organization of the nuclear envelope. Trends Cell
Biol. 4, 127–131.
DAVIES , E., and B. A. L ARKINS. 1980. Ribosomes. In: The Biochemistry of Plants, vol. 1, The Plant Cell, pp. 413–435, N. E. Tolbert,
ed. Academic Press, New York.
GIEGÉ, P., and A. BRENNICKE. 2001. From gene to protein in
higher plant mitochondria. C. R. Acad. Sci., Paris, Sci. de la Vie
324, 209–217.
B ARY, A. 1879. Besprechung. K. Prantl. Lehrbuch der Botanik
für mittlere und höhere Lehranstalten. Bot. Ztg. 37, 221–
223.
GOLDSCHMIDT, E. E. 1988. Regulatory aspects of chlorochromoplast interconversions in senescing Citrus fruit peel.
Isr. J. Bot. 37, 123–130.
DELROT, S., R. ATANASSOVA , E. GOMÈS , and P. COUTOS-THÉVENOT.
2001. Plasma membrane transporters: A machinery for
GÖRLICH, D. 1997. Nuclear protein import. Curr. Opin. Cell Biol.
9, 412–419.
DE
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 39
GÖRLICH, D., and I. W. MATTAJ. 1996. Nucleocytoplasmic transport. Science 271, 1513–1518.
GRAY, J. C. 1996. Biogenesis of chloroplasts in higher plants. In:
Membranes: Specialized Functions in Plants, pp. 441–458, M.
Smallwood, J. P. Knox, and D. J. Bowles, eds. BIOS Scientific,
Oxford.
GRAY, J. C., J. A. SULLIVAN, J. M. HIBBERD, and M. R. HANSEN.
2001. Stromules: Mobile protrusions and interconnections
between plastids. Plant Biol. 3, 223–233.
GRAY, M. W. 1989. Origin and evolution of mitochondrial DNA.
Annu. Rev. Cell Biol. 5, 25–50.
GRAY, M. W. 2000. Mitochondrial genes on the move. Nature
408, 302–305.
GRÖNEGRESS , P. 1971. The greening of chromoplasts in Daucus
carota L. Planta 98, 274–278.
GUNNING, B. E. S. 2001. Membrane geometry of “open” prolamellar bodies. Protoplasma 215, 4–15.
GUNNING, B. E. S., and M. SAMMUT. 1990. Rearrangements of
microtubules involved in establishing cell division planes start
immediately after DNA synthesis and are completed just
before mitosis. Plant Cell 2, 1273–1282.
HAGEMANN, R. 1960. Die Plastidenentwicklung in TomatenKotyledonen. Biol. Zentralbl. 79, 393–411.
HALE, K. L., S. P. MCGRATH, E. LOMBI, S. M. STACK, N. TERRY, I.
J. PICKERING , G. N. GEORGE, and E. A. H. PILON -SMITS. 2001.
Molybdenum sequestration in Brassica species. A role for
anthocyanins? Plant Physiol. 126, 1391–1402.
HANSMANN, P., R. JUNKER , H. SAUTER , and P. SITTE. 1987. Chromoplast development in daffodil coronae during anthesis.
J. Plant Physiol. 131, 133–143.
HANSTEIN, J. 1880. Einige Züge aus der Biologie des Protoplasmas.
Botanische Abhandlungen aus dem Gebiet der Morphologie und
Physiologie, Band 4, Heft 2. Marcus, Bonn.
HAUPT, W., and R. SCHEUERLEIN. 1990. Chloroplast movement.
Plant Cell Environ. 13, 595–614.
HAVAUX, M., and K. KLOPPSTECH. 2001. The protective functions
of carotenoid and flavonoid pigments against excess visible
radiation at chilling temperature investigated in Arabidopsis
npq and tt mutants. Planta 213, 953–966.
HAWES , C., C. M. SAINT-JORE, F. BRANDIZZI, H. ZHENG, A. V.
ANDREEVA , and P. BOEVINK. 2001. Cytoplasmic illuminations:
in planta targeting of fluorescent proteins to cellular organelles. Protoplasma 215, 77–88.
HEESE-PECK, A., and N. V. RAIKHEL . 1998. The nuclear pore
complex. Plant Mol. Biol. 38, 145–162.
HEMERLY, A. S., P. C. G. FERREIRA , M. VAN MONTAGU, and D. INZÉ.
1999. Cell cycle control and plant morphogenesis: Is there an
essential link? BioEssays 21, 29–37.
HERMAN, E. M., X. LI, R. T. SU, P. L ARSEN, H.-T. HSU, and H. SZE.
1994. Vacuolar-type H + -ATPases are associated with the
endoplasmic reticulum and provacuoles of root tip cells. Plant
Physiol. 106, 1313–1324.
HICKS , G. R., and N. V. RAIKHEL . 1995. Protein import into the
nucleus: An integrated view. Annu. Rev. Cell Dev. Biol. 11,
155–188.
HOLTZMAN, E. 1992. Intracellular targeting and sorting. BioScience 42, 608–620.
HÖRTENSTEINER , S., E. MARTINOIA , and N. AMRHEIN. 1992. Reappearance of hydrolytic activities and tonoplast proteins in the
regenerated vacuole of evacuolated protoplasts. Planta 187,
113–121.
HU, J., M. A GUIRRE, C. PETO, J. ALONSO, J. ECKER , and J. CHORY.
2002. A role for peroxisomes in photomorphogenesis and
development of Arabidopsis. Science 297, 405–409.
HUNTLEY, R. P., and J. A. H. MURRAY. 1999. The plant cell cycle.
Curr. Opin. Plant Biol. 2, 440–446.
IVANOVA , M., and T. L. ROST. 1998. Cytokinins and the plant cell
cycle: Problems and pitfalls of proving their function. In: Plant
Cell Proliferation and Its Regulation in Growth and Development,
pp. 45–57, J. A. Bryant and D. Chiatante, eds. Wiley, New
York.
JACOBSON, K., E. D. SHEETS , and R. SIMSON. 1995. Revisiting the
fluid mosaic model of membranes. Science 268, 1441–1442.
JACQMARD, A., C. HOUSSA , and G. BERNIER. 1994. Regulation of
the cell cycle by cytokinins. In: Cytokinins: Chemistry, Activity,
and Function, pp. 197–215, D. W. S. Mok and M. C. Mok, eds.
CRC Press, Boca Raton, FL.
JAVOT, H., and C. MAUREL . 2002. The role of aquaporins in root
water uptake. Ann. Bot. 90, 301–313.
JEDD, G., and N.-H. CHUA . 2002. Visualization of peroxisomes
in living plant cells reveals acto-myosin-dependent cytoplasmic streaming and peroxisome budding. Plant Cell Physiol. 43,
384–392.
JOHNSON, T. L., and L. J. OLSEN. 2001. Building new models for
peroxisome biogenesis. Plant Physiol. 127, 731–739.
JONES , A. 2000. Does the plant mitochondrion integrate cellular
stress and regulate programmed cell death? Trends Plant Sci.
5, 225–230.
JUKES , T. H., and S. OSAWA . 1990. The genetic code in mitochondria and chloroplasts. Experientia 46, 1117–1126.
KAGAWA , T., and M. WADA . 2000. Blue light-induced chloroplast
relocation in Arabidopsis thaliana as analyzed by microbeam
irradiation. Plant Cell Physiol. 41, 84–93.
KAGAWA , T., and M. WADA . 2002. Blue light-induced chloroplast
relocation. Plant Cell Physiol. 43, 367–371.
HEPLER , P. K., and B. E. S. GUNNING. 1998. Confocal fluorescence
microscopy of plant cells. Protoplasma 201, 121–157.
KAPLAN, D. R. 1992. The relationship of cells to organisms in
plants: Problem and implications of an organismal perspective. Int. J. Plant Sci. 153, S28–S37.
HERMAN, E. M., and B. A. L ARKINS. 1999. Protein storage bodies
and vacuoles. Plant Cell 11, 601–614.
KAPLAN, D. R., and W. HAGEMANN. 1991. The relationship of cell
and organism in vascular plants. BioScience 41, 693–703.
40 | Esau’s Plant Anatomy, Third Edition
KEEGSTRA , K., and K. CLINE. 1999. Protein import and routing
systems of chloroplasts. Plant Cell 11, 557–570.
KINDL , H. 1992. Plant peroxisomes: Recent studies on function
and biosynthesis. Cell Biochem. Funct. 10, 153–158.
KIRK, J. T. O., and R. A. E. TILNEY-B ASSETT. 1978. The Plastids.
Their Chemistry, Structure, Growth, and Inheritance, rev. 2nd ed.
Elsevier/North-Holland Biomedical Press, Amsterdam.
KJELLBOM, P., C. L ARSSON, I. JOHANSSON, M. KARLSSON, and U.
JOHANSON. 1999. Aquaporins and water homeostasis in
plants. Trends Plant Sci. 4, 308–314.
genesis. In: Control of Energy Metabolism, pp. 245–248, B.
Chance, R. W. Estabrook, and J. R. Williamson, eds. Academic Press, New York.
L ARKINS , B. A., B. P. DILKES , R. A. DANTE, C. M. COELHO, Y.-M.
WOO, and Y. LIU. 2001. Investigating the hows and whys of
DNA endoreduplication. J. Exp. Bot. 52, 183–192.
LEAVER , C. J., and M. W. GRAY. 1982. Mitochondrial genome
organization and expression in higher plants. Annu. Rev. Plant
Physiol. 33, 373–402.
LEE, D. W., and K. S. GOULD. 2002. Why leaves turn red. Am.
Sci. 90, 524–531.
KLUGE, M., A. FISCHER , and I. C. BUCHANAN -BOLLIG. 1982. Metabolic control of CAM. In: Crassulacean Acid Metabolism, pp.
31–50, I. P. Ting and M. Gibbs, eds. American Society of Plant
Physiologists, Rockville, MD.
LEE, J.-Y., B.-C. YOO, and W. J. LUCAS. 2000. Parallels between
nuclear-pore and plasmodesmal trafficking of information
molecules. Planta 210, 177–187.
KNOTH, R., P. HANSMANN, and P. SITTE. 1986. Chromoplasts of
Palisota barteri, and the molecular structure of chromoplast
tubules. Planta 168, 167–174.
LEONARD, R. T., and T. K. HODGES . 1980. The plasma membrane.
In: The Biochemistry of Plants, vol. 1, The Plant Cell, pp. 163–182,
N. E. Tolbert, ed. Academic Press, New York.
KÖHLER , R. H., and M. R. HANSON. 2000. Plastid tubules of
higher plants are tissue-specific and developmentally regulated. J. Cell Sci. 113, 81–89.
LIAN, H.-L., X. YU, Q. YE, X.-S. DING, Y. KITAGAWA , S.-S. KWAK,
W.-A. SU, and Z.-C. TANG. 2004. The role of aquaporin
RWC3 in drought avoidance in rice. Plant Cell Physiol. 45,
481–489.
KÖHLER , R. H., J. CAO, W. R. ZIPFEL , W. W. WEBB , and M. R.
HANSON. 1997. Exchange of protein molecules through connections between higher plant plastids. Science 276, 2039–
2042.
KONDOROSI, E., F. ROUDIER , and E. GENDREAU. 2000. Plant cellsize control: Growing by ploidy? Curr. Opin. Plant Biol. 3,
488–492.
KOST, B., and N.-H. CHUA . 2002. The plant cytoskeleton: Vacuoles and cell walls make the difference. Cell 108, 9–12.
KOZLOWSKI, T. T., and S. G. PALLARDY. 1997. Physiology of Woody
Plants, 2nd ed. Academic Press, San Diego.
KUNAU, W.-H., and R. ERDMANN. 1998. Peroxisome biogenesis:
Back to the endoplasmic reticulum? Curr. Biol. 8,
R299–R302.
KUNTZ, M., J.-L. EVRARD, A. D’HARLINGUE, J.-H. WEIL , and B.
CAMARA . 1989. Expression of plastid and nuclear genes during
chromoplast differentiation in bell pepper (Capsicum annuum)
and sunflower (Helianthus annuus). Mol. Gen. Genet. 216,
156–163.
KUROIWA , H., T. MORI, M. TAKAHARA , S.-Y. MIYAGISHIMA , and T.
KUROIWA . 2002. Chloroplast division machinery as revealed
by immunofluorescence and electron microscopy. Planta 215,
185–190.
KWOK, E. Y., and M. R. HANSON. 2004. Stromules and the
dynamic nature of plastid morphology. J. Microsc. 214,
124–137.
L AKE, J. A. 1981. The ribosome. Sci. Am. 245 (August), 84–97.
L AM, E., D. PONTIER , and O. DEL POZO. 1999. Die and let live—
Programmed cell death in plants. Curr. Opin. Plant Biol. 2,
502–507.
L ARDY, H. A. 1965. On the direction of pyridine nucleotide
oxidation-reduction reactions in gluconeogenesis and lipo-
LJUBEŠIĆ, N., M. WRISCHER , and Z. DEVIDÉ. 1996. Chromoplast
structures in Thunbergia flowers. Protoplasma 193, 174–180.
LOGAN, H., M. B ASSET, A.-A. VÉRY, and H. SENTENAC . 1997.
Plasma membrane transport systems in higher plants: From
black boxes to molecular physiology. Physiol. Plant. 100,
1–15.
LUCAS , W. J., B. DING, and C. VAN DER SCHOOT. 1993. Plasmodesmata and the supracellular nature of plants. New Phytol.
125, 435–476.
MACKENZIE, S., and L. MCINTOSH. 1999. Higher plant mitochondria. Plant Cell 11, 571–586.
MADIGAN, M. T., J. M. MARTINKO, and J. PARKER . 2003. Brock
Biology of Microorganisms, 10th ed. Pearson Education, Upper
Saddle River, NJ.
MAESHIMA , M. 2001. Tonoplast transporters: organization and
function. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52, 469–497.
MANO, S., C. NAKAMORI, M. HAYASHI, A. KATO, M. KONDO, and
M. NISHIMURA . 2002. Distribution and characterization of
peroxisomes in Arabidopsis by visualization with GFP: Dynamic
morphology and actin-dependent movement. Plant Cell Physiol.
43, 331–341.
MARANO, M. R., and N. CARRILLO. 1991. Chromoplast formation
during tomato fruit ripening. No evidence for plastid DNA
methylation. Plant Mol. Biol. 16, 11–19.
MARANO, M. R., and N. CARRILLO. 1992. Constitutive transcription and stable RNA accumulation in plastids during the conversion of chloroplasts to chromoplasts in ripening tomato
fruits. Plant Physiol. 100, 1103–1113.
MARIENFELD, J., M. UNSELD, and A. BRENNICKE. 1999. The
mitochondrial genome of Arabidopsis is composed of both native
and immigrant information. Trends Plant Sci. 4, 495–502.
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 41
MARTÍN, M., S. MORENO DÍAZ DE LA ESPINA , L. F. JIMÉNEZ-GARCÍA ,
M. E. FERNÁNDEZ-GÓMEZ, and F. J. MEDINA . 1992. Further
investigations on the functional role of two nuclear bodies in
onion cells. Protoplasma 167, 175–182.
MARTIN, W. 1999. A briefly argued case that mitochondria and
plastids are descendants of endosymbionts, but that the
nuclear compartment is not. Proc. R. Soc. Lond. B 266,
1387–1395.
MARTINOIA , E. 1992. Transport processes in vacuoles of higher
plants. Bot. Acta 105, 232–245.
MARTY, F. 1999. Plant vacuoles. Plant Cell 11, 587–599.
MATHUR , J., N. MATHUR , and M. HÜLSKAMP. 2002. Simultaneous
visualization of peroxisomes and cytoskeletal elements reveals
actin and not microtubule-based peroxisomal motility in
plants. Plant Physiol. 128, 1031–1045.
NAKAZONO, M., and A. HIRAI. 1993. Identification of the entire
set of transferred chloroplast DNA sequences in the mitochondrial genome of rice. Mol. Gen. Genet. 236, 341–346.
NISHIMURA , M., Y. TAKEUCHI, L. DE BELLIS , and I. HARANISHIMURA . 1993. Leaf peroxisomes are directly transformed
to glyoxysomes during senescence of pumpkin cotyledons.
Protoplasma 175, 131–137.
NISHIMURA , M., M. HAYASHI, K. TORIYAMA , A. KATO, S. MANO,
K. YAMAGUCHI, M. KONDO, and H. HAYASHI. 1998. Microbody
defective mutants of Arabidopsis. J. Plant Res. 111, 329–332.
NIYOGI, K. K. 2000. Safety valves for photosynthesis. Curr. Opin.
Plant Biol. 3, 455–460.
NUGENT, J. M., and J. D. PALMER . 1988. Location, identity, amount
and serial entry of chloroplast DNA sequences in crucifer
mitochondrial DNAs. Curr. Genet. 14, 501–509.
MATILE, P. 1982. Vacuoles come of age. Physiol. Vég. 20,
303–310.
OLSEN, L. J. 1998. The surprising complexity of peroxisome biogenesis. Plant Mol. Biol. 38, 163–189.
MCFADDEN, G. I. 1999. Endosymbiosis and evolution of the plant
cell. Curr. Opin. Plant Biol. 2, 513–519.
OROSS , J. W., and J. V. POSSINGHAM. 1989. Ultrastructural features of the constricted region of dividing plastids. Protoplasma 150, 131–138.
MIERNYK, J. 1989. Leucoplast isolation. In: Physiology, Biochemistry,
and Genetics of Nongreen Plastids, pp. 15–23, C. D. Boyer, J. C.
Shannon, and R. C. Hardison, eds. American Society of Plant
Physiologists, Rockville, MD.
MILLER , E. A., and M. A. ANDERSON. 1999. Uncoating the mechanisms of vacuolar protein transport. Trends Plant Sci. 4,
46–48.
MIRONOV, V., L. DE VEYLDER , M. VAN MONTAGU, and D. INZÉ.
1999. Cyclin-dependent kinases and cell division in plants—
The nexus. Plant Cell 11, 509–521.
MIYAGISHIMA , S.-Y., M. TAKAHARA , T. MORI, H. KUROIWA , T.
HIGASHIYAMA , and T. KUROIWA . 2001. Plastid division is driven
by a complex mechanism that involves differential transition
of the bacterial and eukaryotic division rings. Plant Cell 13,
2257–2268.
MOHR , W. P. 1979. Pigment bodies in fruits of crimson and high
pigment lines of tomatoes. Ann. Bot. 44, 427–434.
MØLLER , I. M. 2001. Plant mitochondria and oxidative stress:
Electron transport, NADPH turnover, and metabolism of
reactive oxygen species. Annu. Rev. Plant Physiol. Plant Mol. Biol.
52, 561–591.
OSTERYOUNG, K. W. 2000. Organelle fission. Crossing the evolutionary divide. Plant Physiol. 123, 1213–1216.
OSTERYOUNG, K. W., and R. S. MCANDREW. 2001. The plastid
division machine. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52,
315–333.
OSTERYOUNG, K. W., and K. A. P YKE. 1998. Plastid division: Evidence for a prokaryotically derived mechanism. Curr. Opin.
Plant Biol. 1, 475–479.
PALMER , J. D., and C. F. DELWICHE. 1998. The origin and evolution
of plastids and their genomes. In: Molecular Systematics of
Plants. II. DNA Sequencing, pp. 375–409, D. E. Soltis, P. S. Soltis,
and J. J. Doyle, eds. Kluwer Academic, Norwell, MA.
PALMGREN, M. G. 2001. Plant plasma membrane H + -ATPases:
Powerhouses for nutrient uptake. Annu. Rev. Plant Physiol. Plant
Mol. Biol. 52, 817–845.
PARIS , N., C. M. STANLEY, R. L. JONES , and J. C. ROGERS. 1996.
Plant cells contain two functionally distinct vacuolar compartments. Cell 85, 563–572.
MULLEN, R. T., C. R. FLYNN, and R. N. TRELEASE. 2001. How are
peroxisomes formed? The role of the endoplasmic reticulum
and peroxins. Trends Plant Sci. 6, 256–261.
PARK, H., S. S. KREUNEN, A. J. CUTTRISS , D. DELLAPENNA , and
B. J. POGSON. 2002. Identification of the carotenoid isomerase provides insight into carotenoid biosynthesis, prolamellar
body formation, and photomorphogenesis. Plant Cell 14,
321–332.
MULLET, J. E. 1988. Chloroplast development and gene expression. Annu. Rev. Plant Physiol. Plant Mol. Biol. 39, 475–502.
PEARCY, R. W. 1990. Sunflecks and photosynthesis in plant canopies. Annu. Rev. Plant Physiol. Plant Mol. Biol. 41, 421–453.
NAKAMURA , K., and K. MATSUOKA . 1993. Protein targeting to
the vacuole in plant cells. Plant Physiol. 101, 1–5.
PIATTELLI, M. 1981. The betalains: Structure, biosynthesis, and
chemical taxonomy. In: The Biochemistry of Plants, vol. 7, Secondary Plant Products, pp. 557–575, E. E. Conn, ed. Academic
Press, New York.
NAKAMURA , S., T. IKEHARA , H. UCHIDA , T. SUZUKI, T.
SODMERGEN. 1992. Fluorescence microscopy of plastid nucleoids and a survey of nuclease C in higher plants with respect
to mode of plastid inheritance. Protoplasma 169, 68–74.
POSSINGHAM, J. V., and M. E. L AWRENCE. 1983. Controls to plastid
division. Int. Rev. Cytol. 84, 1–56.
42 | Esau’s Plant Anatomy, Third Edition
POTUSCHAK, T., and P. DOERNER . 2001. Cell cycle controls:
Genome-wide analysis in Arabidopsis. Curr. Opin. Plant Biol. 4,
501–506.
PRICE, H. J. 1988. DNA content variation among higher plants.
Ann. Mo. Bot. Gard. 75, 1248–1257.
PURDUE, P. E., and P. B. L AZAROW. 2001. Peroxisome biogenesis.
Annu. Rev. Cell Dev. Biol. 17, 701–752.
P YKE, K. A., and C. A. HOWELLS. 2002. Plastid and stromule
morphogenesis in tomato. Ann. Bot. 90, 559–566.
RATAJCZAK, R., G. HINZ, and D. G. ROBINSON. 1999. Localization
of pyrophosphatase in membranes of cauliflower inflorescence cells. Planta 208, 205–211.
RAVEN, P. R., R. F. EVERT, and S. E. EICHHORN. 1992. Biology of
Plants, 5th ed. Worth, New York.
RAVEN, P. R., R. F. EVERT, and S. E. EICHHORN. 2005. Biology of
Plants, 7th ed. Freeman, New York.
REICHELT, S., and J. KENDRICK-JONES. 2000. Myosins. In: Actin: A
Dynamic Framework for Multiple Plant Cell Functions, pp. 29–44,
C. J. Staiger, F. Baluška, D. Volkmann, and P. W. Barlow, eds.
Kluwer Academic, Dordrecht.
REUMANN, S., and K. KEEGSTRA . 1999. The endosymbiotic origin
of the protein import machinery of chloroplastic envelope
membranes. Trends Plant Sci. 4, 302–307.
REUZEAU, C., J. G. MCNALLY, and B. G. PICKARD. 1997. The
endomembrane sheath: A key structure for understanding
the plant cell? Protoplasma 200, 1–9.
ROBERTSON, J. D. 1962. The membrane of the living cell. Sci. Am.
206 (April), 64–72.
ROBINSON, D. G. 1985. Plant membranes. Endo- and plasma
membranes of plant cells. Wiley, New York.
ROBINSON, D. G., and H. DEPTA . 1988. Coated vesicles. Annu.
Rev. Plant Physiol. Plant Mol. Biol. 39, 53–99.
RODERMEL , S. 2001. Pathways of plastid-to-nucleus signaling.
Trends Plant Sci. 6, 471–478.
ROSE, A., S. PATEL , and I. MEIER . 2004. The plant nuclear envelope. Planta 218, 327–336.
SCHLEIDEN, M. J. 1838. Beiträge zur Phytogenesis. Arch. Anat.
Physiol. Wiss. Med. (Müller’s Arch.) 5, 137–176.
SCHUSTER , W., and A. BRENNICKE. 1987. Plastid, nuclear and
reverse transcriptase sequences in the mitochondrial genome
of Oenothera: Is genetic information transferred between
organelles via RNA? EMBO J. 6, 2857–2863.
SCHWANN, TH. 1839. Mikroskopische Untersuchungen über die
Übereinstimmung in der Struktur und dem Wachstum der Thiere
und Pflanzen. Wilhelm Engelmann, Leipzig.
SEUFFERHELD, M., M. C. F. VIEIRA , F. A. RUIZ, C. O. RODRIGUES,
S. N. J. MORENO, and R. DOCAMPO. 2003. Identification of
organelles in bacteria similar to acidocalcisomes of unicellular
eukaryotes. J. Biol. Chem. 278, 29971–29978.
SHANNON, J. C. 1989. Aqueous and nonaqueous methods for
amyloplast isolation. In: Physiology, Biochemistry, and Genetics
of Nongreen Plastids, pp. 37–48, C. D. Boyer, J. C. Shannon,
and R. C. Hardison, eds. American Society of Plant Physiologists, Rockville, MD.
SIEFRITZ, F., B. OTTO, G. P. BIENERT, A. VAN DER KROL , and
R. KALDENHOFF. 2004. The plasma membrane aquaporin
NtAQP1 is a key component of the leaf unfolding mechanism
in tobacco. Plant J. 37, 147–155.
SINGER , S. J., and G. L. NICOLSON. 1972. The fluid mosaic model
of the structure of cell membranes. Science 175, 720–731.
SITTE, P. 1992. A modern concept of the “cell theory.” A perspective on competing hypotheses of structure. Int. J. Plant Sci.
153, S1–S6.
SITTE, P., H. FALK, and B. LIEDVOGEL . 1980. Chromoplasts. In:
Pigments in Plants, 2nd ed., pp. 117–148, F.-C. Czygan, ed.
Gustav Fischer Verlag, Stuttgart.
SMEEKENS , S., P. WEISBEEK, and C. ROBINSON. 1990. Protein transport into and within chloroplasts. Trends Biochem. Sci. 15,
73–76.
SMITH, J. A. 1987. Vacuolar accumulation of organic acids and
their anions in CAM plants. In: Plant Vacuoles: Their Importance
in Solute Compartmentation in Cells and Their Applications in
Plant Biotechnology, pp. 79–87, B. Marin, ed. Plenum Press,
New York.
RUJAN, T., and W. MARTIN. 2001. How many genes in Arabidopsis
come from cyanobacteria? An estimate from 386 protein
phylogenies. Trends Genet. 17, 113–120.
STALS , H., and D. INZÉ. 2001. When plant cells decide to divide.
Trends Plant Sci. 6, 359–364.
SACK, F. D. 1997. Plastids and gravitropic sensing. Planta 203
(suppl. 1), S63–S68.
STEYN, W. J., S. J. E. WAND, D. M. HOLCROFT, and G. JACOBS.
2002. Anthocyanins in vegetative tissues: A proposed unified
function in photoprotection. New Phytol. 155, 349–361.
SACK, F. D., and J. Z. KISS. 1989. Plastids and gravity perception.
In: Physiology, Biochemistry, and Genetics of Nongreen Plastids,
pp. 171–181, C. D. Boyer, J. C. Shannon, and R. C. Hardison,
eds. American Society of Plant Physiologists, Rockville, MD.
SCHÄFFNER, A. R. 1998. Aquaporin function, structure, and
expression: Are there more surprises to surface in water
relations? Planta 204, 131–139.
SCHEER , U., M. THIRY, and G. GOESSENS. 1993. Structure, function
and assembly of the nucleolus. Trends Cell Biol. 3, 236–241.
STOEBE, B., and U.-G. M AIER . 2002. One, two, three: Nature’s
tool box for building plastids. Protoplasma 219, 123–130.
STRANGE, C. 1992. Cell cycle advances. BioScience 42, 252–256.
SUGIURA , M. 1989. The chloroplast chromosomes in land plants.
Annu. Rev. Cell Biol. 5, 51–70.
SZE, H., X. LI, and M. G. PALMGREN. 1999. Energization of plant
cell membranes by H + -pumping ATPases: Regulation and biosynthesis. Plant Cell 11, 677–690.
The Protoplast: Plasma Membrane, Nucleus, and Cytoplasmic Organelles | 43
TAIZ, L., and E. ZEIGER. 2002. Plant Physiology, 3rd ed. Sinauer
Associates, Sunderland, MA.
TALCOTT, B., and M. S. MOORE. 1999. Getting across the nuclear
pore complex. Trends Cell Biol. 9, 312–318.
THEG, S. M., and S. V. SCOTT. 1993. Protein import into chloroplasts. Trends Cell Biol. 3, 186–190.
TITORENKO, V. I., and R. A. RACHUBINSKI. 1998. The endoplasmic
reticulum plays an essential role in peroxisome biogenesis.
Trends Biochem. Sci. 23, 231–233.
TOLBERT, N. E. 1980. Microbodies—Peroxisomes and glyoxysomes. In: The Biochemistry of Plants, vol. 1, The Plant Cell, pp.
359–388, N. E. Tolbert, ed. Academic Press, New York.
TOLBERT, N. E., and E. ESSNER . 1981. Microbodies: Peroxisomes
and glyoxysomes. J. Cell Biol. 91 (suppl. 3), 271s–283s.
TOURNAIRE-ROUX, C., M. SUTKA , H. JAVOT, E. GOUT, P. GERBEAU,
D.-T. LUU, R. BLIGNY, and C. MAUREL . 2003. Cytosolic pH
regulates root water transport during anoxic stress through
gating of aquaporins. Nature 425, 393–397.
TRELEASE, R. N. 1984. Biogenesis of glyoxysomes. Annu. Rev. Plant
Physiol. 35, 321–347.
TROJAN, A., and H. GABRYŚ . 1996. Chloroplast distribution in
Arabidopsis thaliana (L.) depends on light conditions during
growth. Plant Physiol. 111, 419–425.
TSE, Y. C., B. MO, S. HILLMER , M. ZHAO, S. W. LO, D. G.
ROBINSON, and L. JIANG. 2004. Identification of multivesicular
bodies as prevacuolar compartments in Nicotiana tabacum
BY-2 cells. Plant Cell 16, 672–693.
VAN GESTEL , K., R. H. KÖHLER , and J.-P. VERBELEN. 2002. Plant
mitochondria move on F-actin, but their positioning in the
cortical cytoplasm depends on both F-actin and microtubules.
J. Exp. Bot. 53, 659–667.
VIRCHOW, R. 1858. Die Cellularpathologie in ihrer Begründung auf
physiologische und pathologische Gewebelehre. A. Hirschwald,
Berlin.
VISHNEVETSKY, M., M. OVADIS , and A. VAINSTEIN. 1999. Carotenoid sequestration in plants: The role of carotenoid-associated proteins. Trends Plant Sci. 4, 232–235.
WATERS , M. T., R. G. FRAY, and K. A. P YKE. 2004. Stromule formation is dependent upon plastid size, plastid differentiation
status and the density of plastids within the cell. Plant J. 39,
655–667.
WERGIN, W. P., P. J. GRUBER , and E. H. NEWCOMB . 1970. Fine
structural investigation of nuclear inclusions in plants. J.
Ultrastruct. Res. 30, 533–557.
WIEBE, H. H. 1978. The significance of plant vacuoles. BioScience
28, 327–331.
WILLIAMS , W. E., H. L. GORTON, and S. M. WITIAK. 2003. Chloroplast movements in the field. Plant Cell Environ. 26,
2005–2014.
WINK, M. 1993. The plant vacuole: A multifunctional compartment. J. Exp. Bot. 44 (suppl.), 231–246.
XU, Y., and M. R. HANSON. 2000. Programmed cell death during
pollination-induced petal senescence in Petunia. Plant Physiol.
122, 1323–1334.
YAO, N., B. J. EISFELDER , J. MARVIN, and J. T. GREENBERG. 2004.
The mitochondrion—An organelle commonly involved in
programmed cell death in Arabidopsis thaliana. Plant J. 40,
596–610.
YATSUHASHI, H. 1996. Photoregulation systems for light-oriented
chloroplast movement J. Plant Res. 109, 139–146.
YOUNG, T. E., and D. R. GALLIE. 2000. Regulation of programmed
cell death in maize endosperm by abscisic acid. Plant Mol. Biol.
42, 397–414.
YU, X.-H., T. D. PERDUE, Y. M. HEIMER , and A. M. JONES. 2002.
Mitochondrial involvement in tracheary element programmed
cell death. Cell Death Differ. 9, 189–198.
ZIEGLER , H., E. SCHÄFER, and M. M. SCHNEIDER. 1983. Some
metabolic changes during chloroplast-chromoplast transition
in Capsicum annuum. Physiol. Vég. 21, 485–494.